Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Article

Metal-Organic Frameworks for the Capture of


Trace Aromatic Volatile Organic Compounds
Lin-Hua Xie, Xiao-Min Liu, Tao
He, Jian-Rong Li
jrli@bjut.edu.cn

HIGHLIGHTS
MOFs with excellent stability,
hydrophobic crystals, and interior
pore surfaces

Pore-size reduction by long


angular ligands and network
interpenetration

A high benzene uptake up to


1.75 mmol cm 3 at 0.12 kPa and
80 C

Single-crystal structure
determination of aromatic
hydrocarbon-loaded phases

Two highly stable metal-organic frameworks (MOFs) with suitable hydrophobic


pores have been constructed, and they show great potential for application in the
capture of trace aromatic volatile organic compounds in ambient air. One of the
MOF adsorbents even retains single crystallinity after the adsorption of aromatic
hydrocarbons together with water, allowing precise structure determination of its
guest-loaded phases. These results could offer important guidance for the
development of new adsorbents for the capture of harmful air pollutants.

Xie et al., Chem 4, 1911–1927


August 9, 2018 ª 2018 Elsevier Inc.
https://doi.org/10.1016/j.chempr.2018.05.017
Article
Metal-Organic Frameworks for the Capture
of Trace Aromatic Volatile
Organic Compounds
Lin-Hua Xie,1 Xiao-Min Liu,1 Tao He,1 and Jian-Rong Li1,2,*

SUMMARY The Bigger Picture


Air pollution, typically ‘‘haze air,’’ is getting worse in the urban areas of many Aromatic hydrocarbons
countries worldwide. The development of advanced materials for the removal contribute an important fraction
of air pollutants is of great significance for protecting public health. We herein of volatile organic compounds
present two promising metal-organic frameworks (MOFs) that show high cap- (VOCs) in the urban atmosphere;
ture capacities for aromatic volatile organic compounds even at low pressure not only are these directly harmful
and high temperature, e.g. benzene uptakes of 1.65 and 0.71 mmol g1 at to human health, but they are also
0.12 kPa and 80 C. Breakthrough experiments verify their excellent perfor- a class of precursors of haze
mance in capturing trace benzene in humid air, also showing them to be compa- particulate matter. Carbon-based
rable with some benchmark materials, including Carboxen 1000, PAF-1, MIL- adsorbents are commonly used in
101(Cr), and so on. Furthermore, the capture performance is well interpreted the capture of aromatic VOCs;
by the single-crystal structures of guest-loaded phases, where distinct adsorp- however, this type of material
tion sites, multiple synergistic host-guest and guest-guest interactions, and suffers from structural amorphism,
adaptive local flexibility are disclosed. which limits fundamental studies
on their structural optimization for
enhanced adsorption
INTRODUCTION
performance. We demonstrate
Many large cities worldwide have been experiencing severe and persistent haze air that rationally designed metal-
pollution. The effects of air pollution on human health and climate change have aroused organic frameworks (MOFs) show
considerable public concern. Although the formation mechanism of regional haze is not advantages in the capture of
yet fully understood, the particulate pollution during haze events is believed to be aromatic VOCs over carbon-
largely driven by the formation of secondary aerosols, including secondary inorganic based adsorbents, such as high
aerosols (SIAs) and secondary organic aerosols (SOAs).1 SIAs, mostly sulfate, nitrate, adsorption capacity and a clear
and ammonium, are usually formed from gaseous precursors, such as sulfur dioxide, structure-performance
ammonia, and nitrogen oxides. The atmospheric photooxidation of volatile organic relationship. Precise single-crystal
compounds (VOCs) yields oxygenated or nitrated products that partition between structures of the MOFs and
the gas and particulate phases, resulting in the formation of SOAs.2 relevant VOC-loaded MOFs
suggest that the high capture
Aromatic hydrocarbons, such as benzene, toluene, ethylbenzene, and xylenes (BTEX), capacity results from the presence
contribute an important fraction of VOCs in the urban atmosphere, mostly from gaso- of small hydrophobic pores, less
line- and diesel-powered vehicle emissions.3 These hydrocarbons are not only rich in mutually interfered hydrophobic
petrol but also widely used in industry as solvents and/or chemical raw materials. and hydrophilic adsorption sites,
Besides forming SOAs as mentioned above, they are also directly harmful to human and local flexibility in their
health. Particularly, benzene is one of the hazardous VOCs with the highest toxicity structures.
and prevalence in air, and it is classified as a group-one human carcinogen by the World
Health Organization and International Agency for Research on Cancer.4 The many sour-
ces of vaporized benzene in daily life include but are not limited to petrol, tobacco
smoke, paints, furniture waxes, adhesives, and detergents. Because of its high toxicity,
benzene is often replaced by the less toxic toluene or other derivatives in some appli-
cations. However, the long-term health effects of benzene derivatives are usually

Chem 4, 1911–1927, August 9, 2018 ª 2018 Elsevier Inc. 1911


neglected regardless of their deleteriousness. Removal of these air pollutants is thus of
high significance for protecting public health.

Advanced porous materials with high performance in the capture of aromatic VOCs
in ambient air are highly desirable for use in air purification and air-quality moni-
toring. Currently, porous materials used in commercial air purifiers for the removal
of VOCs and in air-quality analysis for the sampling of VOCs are mostly carbon-
based adsorbents.5,6 These adsorbents have the advantages of low cost, high
porosity, and stability, but they are all structurally amorphous. The lack of precise
structural information on these adsorbents limits fundamental studies on their struc-
tural optimization for enhanced adsorption performance.

Metal-organic frameworks (MOFs) have received considerable attention over the


last two decades because of their high porosity, structural order, and tailorability,
as well as great potential in many applications.7–12 In the past few years, much atten-
tion has been paid to the removal of toxic pollutants (including ammonia, sulfur di-
oxide, nitrogen oxides, chemical-warfare-agent simulants, and others) from ambient
air with these newly developed materials.13–25 MOFs possess rich organic content,
which is an intrinsic advantage in the adsorption of aromatic VOCs. Moreover, the
structures of MOFs are highly crystalline and ordered. In some cases, MOFs even
retain single crystallinity after guest removal and inclusion.26,27 The precise struc-
tures of the MOFs with adsorbed guests can thus be determined by the single-crystal
X-ray diffraction technique, which can provide important insight for understanding
the mechanism of related adsorption processes.

However, the use of MOFs for the capture of aromatic VOCs in air also presents
some challenges. First, air contains a high content of water vapor. Many MOFs are
not stable in humid air for long-term use.28 Many MOFs with hydrophilic groups
or atoms on their pore surface preferentially adsorb water molecules over aromatic
hydrocarbons. Second, the effective adsorption of trace aromatic hydrocarbons de-
mands high adsorption affinity between the adsorbates and the MOF adsorbents
even in competition with the water vapor co-existing in the air. Third, the aromatic
hydrocarbons commonly do not offer special interactions such as strong hydrogen
bonding and coordination bonding with MOFs; the design of MOFs to this end is
thus difficult. Consequently, the preferred MOFs should be moisture stable and
not highly hydrophilic and have pores of appropriate size and shape.

A versatile strategy for constructing hydrolytically stable MOFs is to utilize thermo-


dynamically stable inorganic secondary building units (SBUs),29 which are extended
into porous frameworks by the bridging of organic ligands through strong coordination
bonds. When long linear ligands are utilized to bridge these rigid SBUs, the resulted
MOFs often contain large channels or cavities, which are not favored here, although
sometimes network interpenetration occurs (Scheme 1A). An alternative approach for
the construction of target MOFs for aromatic VOCs adsorption is to replace each linear
ligand in MOFs with a pair of angular ligands that contain non-coordinated aromatic
bridging backbones, where small and hydrophobic pores are expected to form as a
result of the space segmentation by the angular aromatic entities (Scheme 1B). 1BeijingKey Laboratory for Green Catalysis and
Separation and Department of Chemistry and
Chemical Engineering, College of Environmental
We designed and synthesized two MOFs, [Zr6(m3-O)4(m3-OH)4(BDB)6] (BUT-66) and and Energy Engineering, Beijing University of
[Zr6(m3-O)4(m3-OH)4(NDB)6] (BUT-67), by using long and angular ligands (with a Technology, Beijing 100124, P.R. China
bridging angle of 120 ) 4,40 -(benzene-1,3-diyl)dibenzoate (BDB2) and 4,40 -(naph- 2Lead Contact
thalene-2,7-diyl)dibenzoate (NDB2), respectively. They have two-fold interpene- *Correspondence: jrli@bjut.edu.cn
trated three-dimensional (3D) pcu-type frameworks with small pore size, permanent https://doi.org/10.1016/j.chempr.2018.05.017

1912 Chem 4, 1911–1927, August 9, 2018


angular ligand
interpenetration replacement

A B

MOF

linear ligand angular ligand small pore large pore

Scheme 1. Two Approaches to Constructing MOFs with Shrunken Pores


(A) By network interpenetration.
(B) By replacing linear ligands with angular ligands.

porosity, high hydrolytic stability, hydrophobic pore and crystal surfaces, and excel-
lent capture capacities toward aromatic VOCs. Particularly, BUT-66 represents high
volumetric benzene adsorption capacity even at high temperature and low pressure,
outperforming several benchmark adsorbents: MIL-101(Cr),30 ZIF-8 (also referred to
as MAF-4),31,32 MCM-41,33 PAF-1,34 and commercial carbon molecular sieve Car-
boxen 1000. BUT-66 is also capable of trapping ppm levels of benzene in air even
if moisturized, suggesting great potential for practical application. The high perfor-
mance of BUT-66 in the capture of aromatic VOCs can be well interpreted by the sin-
gle-crystal structures of its guest-loaded phases.

RESULTS
Synthesis and Crystal Structures
As-synthesized BUT-66 (referred to as BUT-66a hereafter) and BUT-67 (referred to as
BUT-67a hereafter) single crystals were obtained from the solvothermal reactions of
ZrOCl2$8H2O and free ligands (H2BDB and H2NDB, respectively). Although the
stoichiometry of the ligands and Zr4+ atoms is 1:1 in their formulae, a ligand-to-metal
ratio higher than 1:1 (typically 2:1) was necessary for the formation of BUT-66a or
BUT-67a in synthesis. The presence of an appropriate amount of acetic acid as a
coordination modulator in the reaction systems was found to be important for
the growth of single crystals suitable for single-crystal X-ray diffraction (SC-XRD)
structural analysis. SC-XRD study revealed that BUT-66a crystallizes in the R-3
space group and has a two-fold interpenetrated 3D framework structure built from
Zr6(m3-O)4(m3-OH)4(CO2)12 (Zr6) clusters and angular BDB2 ligands (Figures 1A and
1B). In its structure, each Zr6 cluster coordinates with 12 crystallographically equivalent
BDB2 ligands, and each BDB2 ligand, together with another crystallographically
equivalent one, bridges two Zr6 clusters (Figure 1C). The angular BDB2 ligand has a
122 bridging angle, deviating slightly from its ideal conformation (120 ). The central
and peripheral phenyl rings of BDB2 are not coplanar with dihedral angels of 37.5
and 41.0 , respectively (Table S1). As a result, two carboxylate groups in each ligand
are dihedral in an 85.5 angle. From the topological point of view, each pair of BDB2
ligands can be regarded as an edge, the 12-coordinated Zr6 clusters can be regarded
as 6-connected vertices, and thus a distorted pcu net is formed by alternate connection
of the two types of building units (Figures 1C and S1). The single pcu net is highly open
(each edge ca. 2 nm), which results in a two-fold interpenetration in the final structure of
BUT-66a (Figures 1D and S1). After interpenetration, the pores are significantly reduced.
There are 1D channels with a diameter of ca. 6.0 Å along the (001) direction, and they are

Chem 4, 1911–1927, August 9, 2018 1913


Figure 1. Construction and Structure of BUT-66
(A) 12-coordinated Zr 6 (m 3 -O) 4 (m 3 -OH) 4 (CO2 ) 12 cluster.
(B) Angular BDB 2 ligand.
(C) 3D pcu network in BUT-66 and its simplified representation (color code: Zr-O polyhedra, sky blue; C, black; O, red).
(D) Two-fold interpenetrated structure of BUT-66.

interconnected by small ca. 4.0 Å windows defined by pairs of angular BDB2 ligands
(Figure S2). In addition, there are small isolated cavities of ca. 2.5 Å between neighboring
Zr6 clusters along the (001) direction. Clearly, both network interpenetration and the
presence of angular ligands contribute to the formation of these pores with small size
and pore walls predominantly constituted by non-coordinated aromatic bridging back-
bones of the ligands. The void of BUT-66a is occupied by disordered guest molecules,
accounting for 48.9% of the unit cell volume as estimated by PLATON.35 The ligand
H2NDB used for the synthesis of BUT-67a is derived from the ligand H2BDB by changing
its central phenyl ring into a naphthyl ring (Figure S3). As expected, BUT-67a is isostruc-
tural to BUT-66a has slightly larger 1D channels (ca. 7 Å), windows (ca. 5.5 Å) between the
1D channels, and isolated cavities of similar size according to the results of structure
determination and analysis (Figure S4). Void space occupied by disordered solvents in
BUT-67a is 51.9% of the unit cell volume.

Porosity
N2 adsorption studies at 77 K were carried out for BUT-66 and BUT-67 to evaluate
their porosity after the as-synthesized samples were guest exchanged with methanol
and subsequently evacuated at 80 C under high vacuum. Typical type I isotherms
were obtained (Figure 2A) and demonstrated high apparent Brunauer-Emmett-
Teller surface areas of 1,096 and 984 m2 g1, Langmuir surface areas of 1,291 and
1,141 m2 g1, and pore volumes of 0.46 and 0.41 cm3 g1. Notably, the single crys-
tallinity of BUT-66a was maintained after guest removal. SC-XRD structure determi-
nation for BUT-66 (guest-free phase) revealed that the unit cell was slightly shrunken
(1.3%) compared with that of BUT-66a. The pore volume (0.45 cm3 g1) of BUT-66
calculated from its single-crystal structure was very close to the observed value
from the N2 adsorption experiment (0.46 cm3 g1), suggesting high purity of the
bulk sample, which was also verified by powdered X-ray diffraction (PXRD) patterns
(Figure 2C). In contrast, single crystals of BUT-67a cracked after they were first
guest exchanged with methanol or dichloromethane (DCM) and subsequently
evacuated at 80 C. It was found that the cracking of the crystals occurred only
in the latter step because the single crystals were well maintained in the solvents
by visual inspection under a microscope and the SC-XRD experiment for the DCM
phase (DCM@BUT-67). The experimentally observed pore volume of BUT-67
(0.41 cm3 g1) was obviously smaller than the calculated value (0.56 cm3 g1)

1914 Chem 4, 1911–1927, August 9, 2018


A B

C D

Figure 2. Porosity, Stability, and Hydrophobicity of BUT-66 and BUT-67


(A) N 2 adsorption isotherms recorded at 77 K for BUT-66 and BUT-67, as well as their samples after
exposure in water for 1 month and 48 hr, respectively, and water vapor adsorption isotherms
recorded at 298 K (P0 = 3.1699 kPa).
(B) Optical image of a drop of water on a bulk sample of BUT-66; inset: the water contact angle of
BUT-66 (top right) and photo of BUT-66 crystals floating on a water surface (bottom left).
(C and D) PXRD patterns of BUT-66 (C) and BUT-67 (D) samples after being treated under different
conditions.

according to the SC-XRD data for BUT-67a, suggesting a possible structural change
during the guest removal process. Detailed analysis of the structural transformation
from BUT-67a to BUT-67 with PXRD patterns was hindered by the fact that the inten-
sities of most PXRD peaks of BUT-67 were very weak (Figures 2D and S5).

Stability and Hydrophobicity


The PXRD results reveal that BUT-66 retained its crystalline structure after being sus-
pended in pure water or aqueous solutions with pH values ranging from 1 to 10 for
1 month (Figure 2C). The BUT-66 crystals floated on the surface of water as discussed
below, and the BUT-66 suspensions were stirred to facilitate the contact with the
MOF samples and the liquids throughout the stability tests. BUT-67 also showed
high chemical stability in pure water and acidic and weakly basic aqueous solutions (Fig-
ure 2D). The high hydrolytic stability of BUT-66 and BUT-67 was further confirmed by N2
adsorption isotherms of their water-treated samples, which almost overlapped those of
the pristine samples (Figure 2A). Thermogravimetric analysis (TGA) curves for BUT-66
and BUT-67 showed only slight weight losses (ca. 9%) below 400 C, which could corre-
spond to the release of water molecules adsorbed from the air and dehydroxylation of
their Zr6 clusters.36 Above 450 C, abrupt decreases in weight were observed, which
could result from the decomposition of their frameworks (Figure S6).

Interestingly, the crystals of BUT-66 and BUT-67 floated on the surface of water and
acidic or weakly basic aqueous solutions throughout the stability tests (Figures 2B

Chem 4, 1911–1927, August 9, 2018 1915


and S7), although the calculated density of BUT-66 (1.06 g cm3) was slightly higher
than that of water (1 g cm3). This implies that exterior crystal surfaces of the MOFs
are not well wettable, and the crystals are not submerged as a result of the forces
arising from the surface tension of water. Static water contact angle measurements
gave contact angles of 142.8 G 2 and 137.9 G 2 for BUT-66 and BUT-67, respec-
tively, suggesting high hydrophobicity of their crystal exterior surface (Figures 2B
[inset] and S8). Dynamic contact angle measurements revealed that BUT-66
and BUT-67 showed advancing angles of 144.9 and 141.7 , receding angles of
140.4 and 135.8 , and contact angle hystereses of 4.5 and 5.9 , respectively
(Figures S9–S11).

High exterior hydrophobicity is not indicative of a high interior hydrophobicity for


porous solids and vice versa. For example, PESD-1 shows a hydrophobic exterior
surface (water contact angle >150 ) but intrinsic hydrophilic interior pore surface.37
ZIF-8, well known for its highly hydrophobic interior pore surface, has a hydrophilic
exterior crystal surface, as indicated by a water contact angle of 56 .38–40 Water va-
por adsorption experiments were thus performed. At room temperature, BUT-66
and BUT-67 gave quasi-linear isotherms such that water uptake gradually increased
as the water vapor pressure increased (Figure 2A). At P/P0 close to 0.9, the water
uptakes of BUT-66 and BUT-67 were 5.7 mmol g1 (10.3 wt %) and 6.9 mmol g1
(12.4 wt %), respectively, much lower than the calculated values (42 and 38 wt %,
respectively) if we assume that their pore volumes were fully occupied by crystallized
water (details of the calculation are provided in the Gas and Vapor Adsorption sec-
tion of the Experimental Procedures). These results thus suggest that interior pore
surfaces of BUT-66 and BUT-67 are moderately hydrophobic.

BTEX Adsorption and Capture


The small pore size, high stability, and hydrophobicity of BUT-66 and BUT-67
prompted us to investigate their adsorption performances for BTEX. Benzene
adsorption studies at 25 C and 80 C were carried out first for BUT-66 and BUT-67.
For comparison, benzene adsorption isotherms were also recorded for two well-
known MOFs (MIL-101(Cr) and ZIF-8), a mesoporous silica (MCM-41), a representa-
tive microporous organic polymer (PAF-1), and a commercial carbon molecular sieve
(Carboxen 1000). MIL-101(Cr) shows ultrahigh porosity and high hydrolytic stability,
and its benzene adsorption capacity largely exceeds that of mesoporous silica SBA-
15, HZSM-5 zeolite, and commercial active carbon.41 ZIF-8 features a highly hydro-
phobic pore system with large cavities (12.5 Å) interconnected by small windows
(3.3 Å).42 MCM-41 contains 1D cylindrical mesopores of about 3–5 nm. PAF-1 is
a diamond-like aromatic framework built from tetraphenylene methane and shows
both high porosity and hydrophobicity.34 Carboxen 1000 is used in packed columns
for gas chromatography for the separation of light hydrocarbons and sampling
VOCs for air analysis.43 Porosities of these adsorbents determined by N2 adsorption
studies at 77 K are summarized in Table S2. Water vapor adsorption isotherms at
25 C for the reference adsorbents were also measured (Figure S33) and confirmed
the high hydrophobicity of ZIF-8 and PAF-1. As shown in Figure S17, the benzene
vapor adsorption isotherms at 25 C revealed that PAF-1, MIL-101(Cr), and MCM-
41 have high gravimetric benzene uptake values at 10 kPa (close to the saturated
vapor pressure of benzene at 20 C) of 20.59, 15.84, and 10.49 mmol g1, respec-
tively, much higher than those of Carboxen 1000 (5.04 mmol g1), ZIF-8
(3.99 mmol g1), BUT-66 (3.97 mmol g1), and BUT-67 (6.14 mmol g1). We also re-
corded adsorption isotherms at 80 C after considering that, in some cases, adsor-
bents need to be used in a hot gas flow to capture harmful VOCs such as
vehicle exhaust emissions. The vapor adsorption isotherms at 80 C revealed that

1916 Chem 4, 1911–1927, August 9, 2018


the gravimetric benzene uptake of BUT-66, BUT-67, Carboxen 1000, and ZIF-8
ranged from 3.48 to 4.02 mmol g1 at 10 kPa, lower than that of MIL-101(Cr)
(6.66 mmol g1) and PAF-1 (6.28 mmol g1) but clearly higher than that of
MCM-41 (1.56 mmol g1) (Figure S18). The maximum gravimetric benzene uptakes
of PAF-1 and MIL-101(Cr) at near room temperature were higher than those of all re-
ported adsorbents (Table S3).

For the capture of low-concentration benzene in air, the adsorption capacity of


the adsorbents at low pressure is important. Among the seven adsorbents,
only BUT-66 and Carboxen 1000 showed isotherms with an abrupt increase
at low pressure (Figures S17–S19). Specifically, at 0.012 kPa and 25 C, BUT-66
and Carboxen 1000 showed uptakes of 2.54 and 2.27 mmol g1 (or
2.70 and 1.07 mmol cm3), respectively, obviously higher than those of ZIF-8
(0.03 mmol g1 or 0.03 mmol cm3), MCM-41 (0.07 mmol g1 or 0.02 mmol cm3),
MIL-101(Cr) (0.62 mmol g1 or 0.27 mmol cm3), PAF-1 (0.73 mmol g1 or 0.23 mmol
cm3), and BUT-67 (1.54 mmol g1 or 1.44 mmol cm3). At 0.12 kPa and 80 C,
BUT-66 and Carboxen 1000 showed uptakes of 1.65 and 1.77 mmol g1 (or 1.75 and
0.83 mmol cm3), respectively, which are again much higher than those of MCM-41
(0.03 mmol g1 or 0.01 mmol cm3), ZIF-8 (0.05 mmol g1 or 0.05 mmol cm3), PAF-1
(0.25 mmol g1 or 0.08 mmol cm3), MIL-101(Cr) (0.44 mmol g1 or 0.19 mmol cm3),
and BUT-67 (0.71 mmol g1 or 0.66 mmol cm3).

The volumetric benzene adsorption capacities of BUT-66 were more than twice as high
as those of Carboxen 1000 at most pressures, although their gravimetric uptakes were
similar (Figure 3A) because the density of BUT-66 (1.06 g cm3) is much higher than that
of Carboxen 1000 (0.47 g cm3). Adsorbents with high volumetric adsorption capacities
are preferred when the volume of a separation column or tank is limited in a real appli-
cation. Moreover, adsorbents with higher densities are also advantageous in the
handling processes of fine-powdered or -shaped materials because of their better
immobility. We also recorded adsorption isotherms at 68 C and 93 C, together with
those at 80 C, for the seven adsorbents to calculate their heats of benzene adsorption
(Qst). Although the isotherms of some adsorbents could not be fitted in full with a typical
adsorption isotherm equation for the Qst calculations, Qst values of the adsorbents were
obtained for partial loading ranges and were still informative for comparing their
host-guest interactions. As shown in Figure 3B, the Qst values of Carboxen 1000
(47.7–61.2 kJ mol1) and BUT-66 (47.8–68.5 kJ mol1) were high even in high loading
ranges of 1.45–3.04 and 2.42–3.02 mmol g1, respectively, indicating their strong inter-
actions with benzene. This result is in line with their type I benzene adsorption isotherms.
Adsorption isotherms of toluene, ethylbenzene, o-xylene, m-xylene, and p-xylene were
also recorded at 80 C for BUT-66, and they all showed type I isotherms with high up-
takes at low pressures (Figure 3C).

Furthermore, in order to evaluate the ability of BUT-66 to capture low-concentration


benzene in ambient air, we carried out gas breakthrough experiments at 25 C and
80 C with a gas mixture (10 ppm benzene in dry or wet air) passing through a column
(quartz tube with a 3 mm outer diameter and 1 mm inner diameter) packed with ca.
10 mg BUT-66 in a gas flow rate of 10 mL min1. Although there is no recommended
safe level of exposure for the concentration of airborne benzene, it is commonly
believed that benzene at ppm level has a negative effect on health. Thus, gas with
a low concentration (10 ppm) of benzene was used for the breakthrough experi-
ments. In addition, ambient air is always humid at different levels, so tests were
also performed after the gas mixtures were humidified to relative humidity (RH) of
50% and 80%. As shown in Figure 3D, for dry gas at 25 C, benzene started to

Chem 4, 1911–1927, August 9, 2018 1917


A MIL-101 ZIF-8 BUT-66 B
4 PAF-1 MCM-41 BUT-67 70
Carboxen 1000
60
C6H6 uptake (mmol cm )
-3

3
50

Qst (kJ mol )


-1
40
2 MIL-101(Cr)
PAF-1
30
Carboxen 1000
1 20 ZIF-8
MCM-41
10 BUT-66
0 BUT-67
0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5
-1
Pressure (kPa) Benzene Loading (mmol g )

C D
3
dry gas
Concentration of C 6H6 (ppb)

200 humid gas (50% RH)


humid gas (80% RH)
Uptake (mmol cm )
-3

150
2

100
toluene
1 ethylbenzene 50
o-xylene
m-xylene
p-xylene 0
0
0.0 0.5 1.0 1.5 0 200 400 600 800 1000 1200 1400
-1
Pressure (kPa) Breakthrough Time (hour g )

Figure 3. Selected Aromatic VOCs Adsorption in BUT-66 and BUT-67


For a Figure360 author presentation of Figure 3, see http//dx.doi:10.1016/j.chempr.2018.05.
017#mmc2.
(A and B) Benzene vapor adsorption isotherms recorded at 80  C (A) and isosteric heat of benzene
adsorption (Q st ) (B) for Carboxen 1000, MIL-101(Cr), ZIF-8, MCM-41, PAF-1, BUT-66, and BUT-67.
(C) Vapor adsorption isotherms of toluene, ethylbenzene, o-xylene, m-xylene, and p-xylene
recorded at 80  C for BUT-66.
(D) Breakthrough curves at 25  C for dry air and humid air (relative humidity [RH] = 50% and 80%,
respectively) containing 10 ppm benzene flowed through a column packed with BUT-66.

penetrate the column after about 10 hr (1,000 hr g1), corresponding to a benzene


capture capacity of ca. 0.27 mmol g1 in the dynamic adsorption process. Repeated
breakthrough tests revealed that the adsorbed benzene molecules in BUT-66 could
be removed at 80 C, and the re-activated sample showed no loss of its benzene cap-
ture capacity (Figure S37). Notably, when the breakthrough experiment was carried
out with a moisturized gas mixture (RH z 50%) with 10 ppm benzene, BUT-66 still
showed an excellent performance for benzene removal. The breakthrough curve
almost overlapped that obtained for dry gas (Figure 3D), suggesting a very slight
decrease for the benzene capture capacity of BUT-66 in the humid gas mixture.
When the humidity of the gas mixture increased to RH z 80%, BUT-66 still showed
70% of its benzene capture capacity in dry conditions (Figure 3D). Benzene started
to penetrate the MOF column obviously earlier when the breakthrough experiments
were carried out at 80 C (Figure S38). Nevertheless, BUT-66 retained a benzene cap-
ture capacity of 0.025 mmol g1 even when the gas mixture was moisturized. The
decrease in benzene capture capacity at 80 C is not unexpected because the
adsorption process is exothermic, and increasing the adsorption temperature nor-
mally suppresses the adsorbate uptake of porous adsorbents. In addition, we also
carried out breakthrough experiments to introduce NO gas (1,000 ppm) into the
gas mixture at 80 C to check whether this common air pollutant has an effect on
the performance of BUT-66 in benzene capture. As shown in Figure S39, the break-
through curves before and after the introduction of NO are similar, suggesting the

1918 Chem 4, 1911–1927, August 9, 2018


feasibility of benzene capture by BUT-66 in the presence of some other competing
gases. For comparison, breakthrough experiments were also performed for
the other adsorbents with a 50% RH gas mixture containing 10 ppm benzene at
25 C. As shown in Figure S40, BUT-66 and Carboxen 1000 captured more
benzene than the other adsorbents with capture capacities of 0.26 mmol g1
(0.28 mmol cm3) and 0.35 mmol g1 (0.16 mmol cm3), respectively. ZIF-8
showed almost no benzene capture capacity at all, although its benzene uptake at
equilibrium from the adsorption isotherms was not the lowest of the seven adsor-
bents tested. The discrepancy originates from its extremely small windows
(3.3 Å) between large cages (12.5 Å), which are kinetically slow to open
and thus not favored for uptake of benzene molecules in a dynamic adsorption pro-
cess. As expected, MCM-41 also showed very low benzene capture capacity
(0.01 mmol g1), and those for BUT-67 (0.15 mmol g1), MIL-101(Cr)
(0.13 mmol g1), and PAF-1 (0.07 mmol g1) were moderate.

To better understand the high adsorption performance of BUT-66 toward trace ben-
zene in humid air, we determined the single-crystal structure for C6H6/H2O-loaded
BUT-66 (C6H6/H2O@BUT-66). We obtained C6H6/H2O@BUT-66 single crystals by
exposing BUT-66 single crystals in a saturated benzene vapor and water vapor
mixture at room temperature for about 10 hr. Structure analysis for C6H6/
H2O@BUT-66 showed that the framework remained nearly unchanged. Two types
of adsorbed water molecules with half occupancy were identified in the pore of
BUT-66. One was located in the isolated cavity between neighboring Zr6 clusters
along the (001) direction (referred to as site iii), and the other was sparsely distrib-
uted along the 1D channel walls (site iv), all hydrogen bonded to carboxylate O
and/or m3-O(H) groups of the Zr6 clusters with O$$$O distances between 3.08 and
3.36 Å (Figure S12). There were also two independent adsorbed benzene molecules
with occupancies of 0.75 and 0.5. One was located inside the window defined by a
pair of BDB2 ligands (site i), and the other was located in the center of the 1D chan-
nel (site ii). Generally, two aromatic rings involved in p-p interactions have three ge-
ometries: parallel-displaced, T-shaped edge-to-face, and eclipsed face-to-face
stacking (Figure S13). The guest benzene molecules adsorbed in BUT-66 all inter-
acted with aromatic rings on the host framework in a T-shaped edge-to-face fashion.
Each benzene molecule interacted with four host phenyl rings simultaneously with
center-center distances between 5.02 and 5.43 Å, the shortest C–H$$$p distances
between 2.97 and 3.43 Å, and dihedral angles between aromatic rings between
53.4 and 81.2 (Figures 4A, 4B, and S14). In addition, relatively weak interactions
were also observed between the two types of guest benzene molecules, as sug-
gested by their center-to-center distances (5.50 and 5.83 Å) and dihedral angles
(73.1 and 54.1 ) (Figure 4C). According to the single-crystal structure of C6H6/
H2O@BUT-66, the theoretical value for saturated benzene uptake of BUT-66 is
3.49 mmol g1 (if we assume that the C6H6 molecules are in full occupancy), corre-
sponding to nine benzene molecules adsorbed per formula (three at site i and six at
site ii), which is close to the maximum uptake (3.48 mmol g1) observed in the vapor
adsorption experiment at 80 C and slightly lower than that at 25 C (3.97 mmol g1)
(Figures S17 and S18), probably because of the extra adsorption on the external sur-
face of the sample. When BUT-66 was exposed at lower partial pressures of benzene
and water, those water and benzene adsorption sites were also observed in the re-
sulting single crystals (C6H6/H2O@BUT-66-2) by SC-XRD, although occupancies of
the guests decreased. The single-crystal structures of BUT-66 loaded with
toluene/H2O, p-xylene/H2O, and m-xylene/H2O were also determined. Similarly,
these hydrocarbon molecules were all located in adsorption sites i and ii, and water
molecules were adsorbed in sites iii and iv.

Chem 4, 1911–1927, August 9, 2018 1919


Figure 4. Single-Crystal X-Ray Diffraction Elucidation of Benzene Adsorption in BUT-66
(A and B) Two types of adsorbed benzene molecules located in the windows between channels (A)
and 1D channels (B).
(C) Distribution of two types of adsorbed benzene molecules in a unit cell of BUT-66.

Clearly, there are specific adsorption sites in BUT-66 for BTEX molecules and water
molecules. Adsorption of non-polar or weakly polar molecules is preferred at sites i
and ii because the pore surface around these sites is constituted by hydrophobic
phenyl groups of the organic ligands. Single-crystal determination for a water-
loaded phase of BUT-66 (H2O@BUT-66) revealed that adsorbed water molecules
sparsely located around sites i and ii, even when the crystal was exposed to high hu-
midity. In contrast, sites iii and iv were close to the Zr6 clusters. Preferential adsorp-
tion of water at the sites was driven by hydrogen-bonding interactions between
water molecules and hydrophilic sites (carboxylate O of the m3-O(H) groups) on
the Zr6 clusters. The adsorption of hydrocarbon molecules on the hydrophobic sites
is thus not evidently affected by the adsorption of water molecules on the hydrophil-
ic sites, which agrees well with the high benzene adsorption capacity of BUT-66 un-
der humid conditions. The high BTEX adsorption capacity of BUT-66 even at high
temperature and low pressure could also result from the multipoint host-guest inter-
actions and adsorption-induced guest-guest interactions. In addition, changes in
the configuration of the ligand BDB2 in BUT-66 were observed after guest inclusion.
As shown in Table S1, the dihedral angles between the two peripheral phenyl rings
and the central phenyl ring in BDB2 obviously vary after BUT-66 adsorbs different
guests, but the 2D geometry of the ligand is nearly retained. Clearly, the peripheral
phenyl rings of BDB2 rotate to optimize the host-guest interactions for the lowest
energy of the system during the adsorption processes. It is difficult to observe this
type of structural flexibility of MOFs in PXRD patterns because the partial rotation
of the ligand without changing its overall 2D geometry does not lead to distinct
changes in the MOF unit cells.

DISCUSSION
In the past few years, zirconium-based MOFs (Zr-MOFs) have been demonstrated to be
a subclass of MOFs with both high stability and good designability, making them an

1920 Chem 4, 1911–1927, August 9, 2018


excellent platform for the development of advanced materials for various applications.
Most Zr-MOFs are based on Zr6 clusters and polytopic aromatic carboxylate ligands.
Because of the bulky size, high connectivity, and rigid coordination geometry of the
Zr6 cluster, Zr-MOFs built from long linear ligands are commonly formed as non-inter-
penetrated 3D frameworks with large cavities and channels. In order to construct
Zr-MOFs with small pores in this work, we chose long angular ligands in place of linear
ligands. We expected that two angular ligands with appropriate geometry would simul-
taneously link two Zr6 clusters to allow a 6-connected net with the pcu topology instead
of 12-connected fcu network to form, although with each Zr6 cluster still coordinated
with 12 ligands. Such a construction would favor the formation of MOFs with small pores
even if long ligands were used. On the one hand, the pcu nets are more prone to inter-
penetrate than fcu nets. On the other hand, small apertures are likely to be generated by
the pairs of angular ligands, where non-coordinated aromatic backbones of the angular
ligands serve as the pore walls. Such small hydrophobic pores are preferred for the
adsorption of small aromatic hydrocarbon molecules. We noticed that several Zr-
MOFs with angular dicarboxylate ligands, including DUT-51,44 DUT-67,45 MOF-802,46
BUT-10, and BUT-11,47 had already been reported in the literature. However, in these
Zr-MOFs, the linker bridging two neighboring Zr6 clusters is a single ligand rather
than a pair of ligands. The angular ligands in these Zr-MOFs are only slightly bent (Fig-
ure S15). For example, BUT-10 and BUT-11 have an fcu-type network with 160 angle
ligands; MOF-802, DUT-51, and DUT-67 are constructed from slightly more bent li-
gands (147 156 ), and they are either a bct-type network (10-connected) or an reo-
type network (8-connected). By using more bent ligands (BDB2 and NDB2 at
120 ), we successfully constructed BUT-66 and BUT-67 with pcu-type networks in
this work. The two MOFs represent rare examples of Zr-MOFs with the pcu topology.48
The angular ligands in BUT-66 or BUT-67 are different from those in any of the
above-mentioned Zr-MOFs by not only the bridging angle but also the 3D spatial dis-
tribution of their carboxylate groups. The two carboxylate groups in the former are
non-coplanar with a dihedral angle of 90 . In contrast, in the latter, they are ideally
coplanar (Figure S15). The above findings highlight the importance of ligand geometry
in the construction of MOFs with desired pore structure and thus application-oriented
properties.

Interestingly, the two-fold interpenetrated BUT-66 and BUT-67 represent high and
moderate hydrophobicity, respectively, in both the exterior crystal surface and the
interior pore surface. Only a few reported MOFs show high exterior hydrophobicity,
and most of them are either constructed from ligands with alkyl chains or fluorine
groups or obtained by post-modification with hydrophobic molecules or polymers
(Table S4 and associated references). In contrast, in our case, the crystal surfaces
of BUT-66 and BUT-67 were not post-modified, and their ligands featured only
angular aromatic backbones without an alkyl or fluorine substituent. Thus, their exte-
rior hydrophobicity is probably related to the nano- and microstructures of their crys-
tal surfaces.37,39,49,50 The moderately hydrophobic interior pore surfaces of BUT-66
and BUT-67 could be interpreted from their single-crystal structures. As mentioned
above, their pore walls are mostly made of aromatic backbones of the angular li-
gands, which are weakly polar. Furthermore, close inspection of the structures indi-
cates that the two-fold interpenetration of the two Zr-MOFs also contributes to their
interior hydrophobicity because hydrophilic carboxylate O atoms of the Zr6 clusters
in one network are partially sheltered by hydrophobic phenyl groups of ligands in the
other network (Figure S16), which leads their pore surfaces to be predominantly
lined with aromatic rings of the ligands. This result thus suggests an alternative
way of constructing hydrophobic MOFs with well-defined inorganic SBUs and long
angular ligands.

Chem 4, 1911–1927, August 9, 2018 1921


We demonstrated that BUT-66 has quite a high uptake for BTEX even at low
pressure and high temperature. Particularly, the benzene uptake of BUT-66 at
0.12 kPa and 80 C is 1.75 mmol cm3, which is higher than those of mesoporous
MCM-41 (0.01 mmol cm3), highly hydrophobic ZIF-8 (0.05 mmol cm3), highly
porous MIL-101(Cr) (0.19 mmol cm3), highly porous and hydrophobic PAF-1
(0.08 mmol cm3), and carbon-based Carboxen 1000 (0.83 mmol cm3). This MOF
is also capable of capturing traces of benzene in air (10 ppm) at room temperature
with a capacity of ca. 0.28 mmol cm3, and the benzene capture capacity of
BUT-66 is nearly uncompromised when the gas mixture is moisturized with RH z
50%. The high BTEX capture performance of BUT-66 can be directly related to its
pore structure, as rationalized by the single-crystal structure analyses of its guest-
loaded phases. The pores in BUT-66 are relatively small and uniform, and the pore
walls are predominantly hydrophobic phenyl rings of the ligands as a result of the
face-to-face distributed angular ligands and network interpenetration. The adjacent
hydrophobic pore walls could induce multiple synergistic host-guest and guest-
guest interactions between the framework and the BTEX molecules, giving rise to
a high driving force for the adsorption of these hydrocarbons in BUT-66. In terms
of single-crystal structures, several well-defined adsorption sites are clearly selective
toward hydrophobic hydrocarbon molecules and hydrophilic water molecules.
Adsorption of any type of molecule on the corresponding adsorption sites would
not affect the adsorption of the others evidently. Consequently, co-adsorption of
water and BTEX molecules in BUT-66 could proceed without significantly lowering
its adsorption capacity for the hydrocarbons. This is important for practical applica-
tion of the material in the removal of hydrocarbons from air, where water always
exists.

We also found that the framework of BUT-66 has local flexibility during the BTEX
adsorption processes, although the unit cell of the overall framework remained
nearly unchanged. The local flexibility manifested by the rotation of peripheral
phenyl rings of the ligands in BUT-66 could indeed maximize the host-guest and
guest-guest interactions to some extent to lower the energy of the system. The local
flexibility also explains why BUT-66 is capable of including all BTEX molecules,
although its pores at equilibrium (4.0–6.0 Å) are small in comparison with the kinetic
diameters of the hydrocarbons (benzene, toluene, and p-xylene, 5.8 Å; ethylben-
zene, 6.0 Å; o-xylene and m-xylene, 6.8 Å).51 Overall, uniform, small, and hydropho-
bic pores and local structural flexibility account for the excellent performance of
BUT-66 in BTEX capture.

In summary, we have synthesized two highly stable MOFs, built from long 120
angular ligands and robust Zr6 clusters, that feature two-fold interpenetrated pcu
network structures and small hydrophobic pores. They show great potential for
the capture of aromatic VOCs in ambient air. Particularly, one of the MOFs,
BUT-66, shows significantly higher benzene adsorption capacity (1.75 mmol cm3)
than some benchmark materials, such as MCM-41, MIL-101(Cr), ZIF-8, PAF-1, and
Carboxen 1000, at low pressure and high temperature. Furthermore, this MOF is
capable of capturing several parts per million of benzene in humid air (RH 50%) in
a high capacity (ca. 0.28 mmol cm3). Single-crystal structure analyses for BUT-66
and its guest-loaded phases suggest that its high performance in the capture of
aromatic VOCs results from the small hydrophobic pores, mutually less-interfered
hydrophobic and hydrophilic adsorption sites, and rigid overall framework with local
rotatable phenyl rings. This work demonstrates the promise of MOFs as new
advanced materials for practical use in cleaning air and suggests an approach for
constructing highly stable MOFs with small and uniform hydrophobic micropores.

1922 Chem 4, 1911–1927, August 9, 2018


EXPERIMENTAL PROCEDURES
Materials
The ligands H2BDB and H2NDB and MOFs MIL-101(Cr), PAF-1, and ZIF-8 were
synthesized according to previously reported methods.34,52–54 Carboxen 1000 was
purchased from Sigma-Aldrich. MCM-41 was purchased from Nanjing XFNANO
Materials Tech Co.

Synthesis of [Zr6(m3-O)4(m3-OH)4(BDB)6] (BUT-66)


In a 5 mL glass vial, ZrOCl2$8H2O (0.0322 g, 0.1 mmol) and H2BDB (0.0636 g,
0.2 mmol) were added in a mixture of 1.5 mL of N,N-dimethylformamide (DMF)
and 0.25 mL of acetic acid. The vial was then sealed and kept under ultrasonication
for about 30 min to dissolve all the starting materials. The mixture was heated at
120 C for 24 hr and then cooled down to room temperature. Colorless hexagonal
crystals were obtained by filtration. After the as-synthesized crystals were guest
exchanged with methanol for 3 days (20 mL 3 3) and subsequently evacuated at
80 C under high vacuum, 0.035 g (yield ca. 81% based on Zr) of the final product
of BUT-66 was obtained. Anal. calcd. (%) for C120H76O32Zr6: C, 55.92; H, 2.97.
Found: C, 55.45; H, 3.06. FT-IR (KBr pellet, cm1): 3412(m), 3068(w), 2927(w),
1662(s), 1583(s), 1412(s), 1257(m), 1153(w), 1087(w), 777(s), 653(s), 457(m).

Synthesis of [Zr6(m3-O)4(m3-OH)4(NDB)6] (BUT-67)


The synthesis of BUT-67 followed that of BUT-66, except that ligand H2BDB was re-
placed with H2NDB; 0.037 g (yield ca. 77% based on Zr) of product BUT-67 was ob-
tained. Anal. calcd. (%) for C144H84O32Zr6: C, 60.19; H, 2.95. Found: C, 59.84; H,
3.11. FT-IR (KBr pellet, cm1): 3421(m), 3057(w), 2918(w), 1652(s), 1602(s), 1543(s),
1408(s), 1180(m), 1020(w), 851(m), 770(s), 659(m), 472(m).

Single-Crystal X-Ray Diffraction


The single-crystal diffraction data for BUT-66a, BUT-66, guest-adsorbed
phases of BUT-66 (C6H6/H2O@BUT-66, C6H6/H2O@BUT-66-2, toluene/H2O@BUT-
66, p-xylene/H2O@BUT-66, m-xylene/H2O@BUT-66, H2O@BUT-66), BUT-67a, and
DCM@BUT-67 were collected with a Rigaku Supernova CCD diffractometer equip-
ped with mirror-monochromatic enhanced Cu-Ka radiation (l = 1.54184 Å) at 100
or 353 K (353 K for BUT-66 only, as shown in Figure S41). The dataset was corrected
by empirical absorption correction with spherical harmonics implemented in the
SCALE3 ABSPACK scaling algorithm.55 The structure was solved by direct methods
and refined by full-matrix least squares on F2 with anisotropic displacement with the
SHELXTL software package.56 Non-hydrogen atoms on the frameworks were refined
with anisotropic displacement parameters during the final cycles. The hydrogen
atoms on the ligands were positioned geometrically and refined with a riding model.
The electron density of the disordered guest molecules in BUT-66a and BUT-67a
were flattened according to the SQUEEZE routine in PLATON.35 Because of the ther-
mal motion effect, the atoms of adsorbed guests show larger atomic displacements
than those on the framework. Thus, geometry constraints were applied to the ad-
sorbed hydrocarbon molecules to maintain their reasonable configurations during
the structure refinements. The crystal data and structure refinement results for these
compounds are listed in Tables S5–S14, and their ORTEP plots of asymmetry units
are shown in Figures S42 and S43. The single crystals of C6H6/H2O@BUT-66 were
prepared according to the following steps: methanol-exchanged BUT-66 single
crystals were heated at 80 C overnight under a nitrogen flow for the removal of all
guests. Then, the guest-free BUT-66 single crystals were transferred to a sealed flask,
which contained two open vials filled with benzene and water, respectively; after re-
maining in the saturated benzene and water vapor atmosphere overnight at room

Chem 4, 1911–1927, August 9, 2018 1923


temperature, the crystals were used for SC-XRD measurements. The single crystals
of H2O@BUT-66, toluene/H2O@BUT-66, p-xylene/H2O@BUT-66, and m-xylene/
H2O@BUT-66 were prepared similarly except that the liquid benzene was replaced
by other hydrocarbons or was not introduced (for H2O@BUT-66). Single crystals of
C6H6/H2O@BUT-66-2 were prepared similarly to C6H6/H2O@BUT-66 except that
the flask containing guest-free BUT-66, benzene, and water was kept at 6 C instead
of 25 C for 24 hr. This experiment was performed for checking how adsorbed ben-
zene and water molecules locate in BUT-66 when their partial pressure is reduced.
Structure determination and refinement revealed that the adsorption sites for ben-
zene and water molecules in C6H6/H2O@BUT-66-2 are the same as those in C6H6/
H2O@BUT-66, although their occupancies decrease. Occupancies of the guests
are listed as follows. C6H6/H2O@BUT-66: C6H6 at site i, 0.75; C6H6 at site ii, 0.5;
H2O, 0.5. C6H6/H2O@BUT-66-2: C6H6 at site i, 0.5; C6H6 at site ii, 0.1; H2O, 0.5.
toluene/H2O@BUT-66: toluene at site i, 0.5; toluene at site ii, 0.25; H2O, 0.75.
p-xylene/H2O@BUT-66: p-xylene at site i, 0.5; p-xylene at site ii, 0.25; H2O, 0.75.
m-xylene/H2O@BUT-66: m-xylene at site i, 0.25; m-xylene at site ii, 0.25; H2O,
0.375; H2O@BUT-66: H2O, 0.125–0.25; DCM@BUT-67: CH2Cl2, 1.

Gas and Vapor Adsorption


N2 and NO gas adsorption isotherms and vapor adsorption isotherms for water,
benzene, toluene, ethylbenzene, and xylenes were carried out with a Micromeritics
ASAP2020 surface area and pore analyzer, which was customized with all stainless-
steel parts inside the manifold (Figures S17–S34 and Table S15–S27). The adsorption
temperatures were controlled by a liquid nitrogen bath (77 K), a water bath (298 K),
or a heating mantle (341, 353, and 366 K). For hydrocarbon vapor adsorption exper-
iments, the plastic stopper of the sample tube was removed during the adsorption
processes to avoid extra adsorption of hydrocarbon vapors by the plastic stopper.
For the adsorption experiments performed at 341, 353, or 366 K, part of the sample
tube was exposed at room temperature. To avoid vapor condensation at the
exposed part, adsorption isotherms were recorded only at pressures much lower
than the saturated vapor pressure of the hydrocarbons at those temperatures. The
calculated maximum water uptakes for BUT-66 and BUT-67 were predicted by the
following semi-empirical equation: uptake in the percentage of weight (wt %) =
rice (g cm3) 3 pore volume (cm3 g1) 3 100%, where rice is the density of crystallized
water (ice, 0.9168 g cm3) at near room temperature; the pore volume was obtained
from N2 adsorption experiments at 77 K. We justified the validity of this calculation
method by applying it to calculate the maximum water uptake of MIL-101(Cr). The
calculated maximum water uptake was 143 wt % (pore volume of MIL-101(Cr),
1.56 cm3 g1; Table S2), and the experimentally observed uptake at P/P0 = 0.94
(near saturation) was 137 wt % (Figure S33).

Estimation of Isosteric Heat of the Benzene Adsorption


Isosteric heat of the benzene adsorption was estimated from the benzene sorption
data measured at 341, 353, and 366 K. First, the Toth equation57 (Equation 1) was
used for independent fitting of the isotherms (Figures S26–S32), where N is the
amount of adsorbed gas in mmol g1, Nsat is the saturated adsorption amount, P
is the pressure in kPa, and b and t are the equation constants. Satisfactory fitting re-
sults were obtained only for partial loading ranges, although some other adsorption
isotherm models were also tested. The Toth equation reduces to the Langmuir equa-
tion when t equals 1.

Nsat bP
N= : (Equation 1)
ð1 + bP t Þ1=t

1924 Chem 4, 1911–1927, August 9, 2018


The expression for the pressure P in terms of the benzene adsorption amount N can
be obtained by Equation 2:

N
P= 1=t : (Equation 2)
bt Ntsat  bNt

Isosteric heat (Qst) of the benzene adsorption was calculated with the Clausius-Cla-
peyron equation58 (Equation 3), where T is the temperature, R is the universal gas
constant, and C is a constant. The Qst values at different benzene loading N were
obtained from the slopes of the plots of (lnP)N as a function of (1/T).

Qst 1
ðln PÞN =  +C (Equation 3)
R T

Breakthrough Experiments
The breakthrough experiments were carried out with a calibration gas (ca. 20
ppm benzene balanced with dry air) by a setup shown in Figure S35. Quartz
tubes (3 mm outer diameter, 1 mm inner diameter, and 50 mm length) were
packed with BUT-66. A test run without adsorbent was also carried out (Fig-
ure S36). For the dry gas breakthrough experiment, a mixture of the calibration
gas and dry air (1:1 v/v) was passed through the packed tube at a flow rate of
10 mL min1. The humid gas with 50% RH breakthrough experiment was similar
to the dry gas breakthrough experiment except that the dry air flow was pre-
saturated with water vapor by being bubbled through deionized water before
being mixed with the calibration gas flow. For experiments with 80% RH humid
gas, a calibration gas containing 50 ppm benzene was used, and the ratio of
calibration gas to dry gas was set to 1:4 (v/v) with a total flow rate of 10 mL
min1. For experiments with NO gas (ca. 1,000 ppm), air was replace with nitro-
gen for preventing NO oxidation. The flow rate of the gases was controlled by a
mass flow controller (Alicat). The concentration of benzene in the gases passing
through the column was monitored with a Hiden HPR20 mass spectrometer gas
analysis system.

DATA AND SOFTWARE AVAILABILITY


The single-crystal X-ray diffraction data of the MOFs reported in this work have been
deposited in the Cambridge Crystallographic Data Centre (CCDC) under accession
numbers CCDC: 1574373 (BUT-66), 1574374 (BUT-66a), 1574375 (C6H6/H2O@BUT-
66), 1574376 (toluene/H2O@BUT-66), 1574377 (p-xylene/H2O@BUT-66), 1574378
(m-xylene/H2O@BUT-66), 1574379 (BUT-67a), 1574380 (DCM@BUT-67), 1838071
(C6H6/H2O@BUT-66-2), and 1838075 (H2O@BUT-66).

SUPPLEMENTAL INFORMATION
Supplemental Information includes 43 figures, 27 tables, and 1 data file and can be
found with this article online at https://doi.org/10.1016/j.chempr.2018.05.017.

ACKNOWLEDGMENTS
This work was financially supported by the Natural Science Foundation of China
(grant nos. 21576006 and 21601008), the Science Fund for Creative Research
Groups of the National Natural Science Foundation of China (grant no.
51621003), the China Postdoctoral Science Foundation (grant nos. 2015M580027
and 2016M600879), and the Beijing Natural Science Foundation (grant no.
2182005).

Chem 4, 1911–1927, August 9, 2018 1925


AUTHOR CONTRIBUTIONS
Conceptualization, J.-R.L. and L.-H.X.; Validation, J.-R.L., L.-H.X., and X.-M.L.;
Formal Analysis, L.-H.X., X.-M.L., and T.H.; Investigation, L.-H.X., X.-M.L., and
T.H.; Resources, J.-R.L.; Writing – Original Draft, J.-R.L. and L.-H.X.; Writing – Review
& Editing, J.-R.L., L.-H.X., X.-M.L., and T.H.; Visualization, L.-H.X.; Supervision,
J.-R.L.; Project Administration, J.-R.L.; Funding Acquisition, J.-R.L., L.-H.X., and
X.-M.L.

DECLARATION OF INTERESTS
The authors declare no competing interests.

Received: September 21, 2017


Revised: November 20, 2017
Accepted: May 24, 2018
Published: June 21, 2018

REFERENCES AND NOTES


1. Huang, R.-J., Zhang, Y., Bozzetti, C., Ho, K.-F., propylene from propane. Science 353, 19. Rodrı́guez-Albelo, L.M., López-Maya, E.,
Cao, J.-J., Han, Y., Daellenbach, K.R., Slowik, 137–140. Hamad, S., Ruiz-Salvador, A.R., Calero, S., and
J.G., Platt, S.M., Canonaco, F., et al. (2014). Navarro, J.A.R. (2017). Selective sulfur dioxide
High secondary aerosol contribution to 11. Sato, H., Kosaka, W., Matsuda, R., Hori, A., adsorption on crystal defect sites on an
particulate pollution during haze events in Hijikata, Y., Belosludov, R.V., Sakaki, S., Takata, isoreticular metal organic framework series.
China. Nature 514, 218–222. M., and Kitagawa, S. (2014). Self-accelerating Nat. Commun. 8, 14457.
CO sorption in a soft nanoporous crystal.
2. Ng, N., Kroll, J., Chan, A., Chhabra, P., Flagan, Science 343, 167–170. 20. Mondloch, J.E., Katz, M.J., Isley Iii, W.C.,
R., and Seinfeld, J. (2007). Secondary organic Ghosh, P., Liao, P., Bury, W., Wagner, G.W.,
aerosol formation from m-xylene, toluene, and 12. Mason, J.A., Oktawiec, J., Taylor, M.K., Hall, M.G., DeCoste, J.B., Peterson, G.W., et al.
benzene. Atmos. Chem. Phys. 7, 3909–3922. Hudson, M.R., Rodriguez, J., Bachman, J.E., (2015). Destruction of chemical warfare agents
Gonzalez, M.I., Cervellino, A., Guagliardi, A., using metal–organic frameworks. Nat. Mater.
3. Gentner, D.R., Jathar, S.H., Gordon, T.D., Brown, C.M., et al. (2015). Methane storage in 14, 512–516.
Bahreini, R., Day, D.A., El Haddad, I., Hayes, flexible metal–organic frameworks with
P.L., Pieber, S.M., Platt, S.M., de Gouw, J., et al. intrinsic thermal management. Nature 527, 21. Zhai, Q.-G., Bu, X., Mao, C., Zhao, X., Daemen,
(2017). Review of urban secondary organic 357–361. L., Cheng, Y., Ramirez-Cuesta, A.J., and Feng,
aerosol formation from gasoline and diesel P. (2016). An ultra-tunable platform for
motor vehicle emissions. Environ. Sci. Technol. 13. Barea, E., Montoro, C., and Navarro, J.A.R. molecular engineering of high-performance
51, 1074–1093. (2014). Toxic gas removal - metal-organic crystalline porous materials. Nat. Commun. 7,
frameworks for the capture and degradation of 13645.
4. WHO Regional Office for Europe. (2010). WHO toxic gases and vapours. Chem. Soc. Rev. 43,
Guidelines for Indoor Air Quality: Selected 5419–5430. 22. Britt, D., Tranchemontagne, D., and Yaghi,
pollutants (WHO). O.M. (2008). Metal-organic frameworks with
14. DeCoste, J.B., and Peterson, G.W. (2014). high capacity and selectivity for harmful gases.
5. Khan, F.I., and Ghoshal, A.K. (2000). Removal of Metal–organic frameworks for air purification Proc. Natl. Acad. Sci. USA 105, 11623–11627.
volatile organic compounds from polluted air. of toxic chemicals. Chem. Rev. 114, 5695–5727.
J. Loss Prevent. Proc. Ind. 13, 527–545. 23. Wisser, D., Wisser, F.M., Raschke, S., Klein, N.,
15. Yu, Y., Zhang, X.-M., Ma, J.-P., Liu, Q.-K., Wang, Leistner, M., Grothe, J., Brunner, E., and Kaskel,
6. Slominska, M., Krol, S., and Namiesnik, J.
P., and Dong, Y.-B. (2014). Cu(I)-MOF: naked- S. (2015). Biological Chitin-MOF composites
(2013). Removal of BTEX compounds from
eye colorimetric sensor for humidity and with hierarchical pore systems for air-filtration
waste gases; destruction and recovery
techniques. Crit. Rev. Environ. Sci. Technol. 43,
formaldehyde in single-crystal-to-single- applications. Angew. Chem. Int. Ed. 54, 12588–
crystal fashion. Chem. Commun. 50, 1444– 12591.
1417–1445.
1446.
7. Furukawa, H., Cordova, K.E., O’Keeffe, M., and 24. Rieth, A.J., Tulchinsky, Y., and Dinca, M. (2016).
Yaghi, O.M. (2013). The chemistry and 16. Jasuja, H., Peterson, G.W., Decoste, J.B., High and reversible ammonia uptake in
applications of metal-organic frameworks. Browe, M.A., and Walton, K.S. (2015). mesoporous azolate metal organic frameworks
Science 341, 974. Evaluation of MOFs for air purification and air with open Mn, Co, and Ni sites. J. Am. Chem.
quality control applications: Ammonia removal Soc. 138, 9401–9404.
8. Kaskel, S. (2016). The Chemistry of Metal- from air. Chem. Eng. Sci. 124, 118–124.
organic Frameworks: Synthesis, 25. Cui, X., Yang, Q., Yang, L., Krishna, R., Zhang,
Characterization, and Applications (John 17. Vellingiri, K., Szulejko, J.E., Kumar, P., Kwon, Z., Bao, Z., Wu, H., Ren, Q., Zhou, W., Chen, B.,
Wiley). E.E., Kim, K.-H., Deep, A., Boukhvalov, D.W., et al. (2017). Ultrahigh and selective SO2 uptake
and Brown, R.J. (2016). Metal organic in inorganic Anion-Pillared hybrid porous
9. Cui, X.L., Chen, K.J., Xing, H.B., Yang, Q.W., frameworks as sorption media for volatile and materials. Adv. Mater. 29, 1606929.
Krishna, R., Bao, Z.B., Wu, H., Zhou, W., Dong, semi-volatile organic compounds at ambient
X.L., Han, Y., et al. (2016). Pore chemistry and conditions. Sci. Rep. 6, 27813. 26. Zhang, J.-P., Liao, P.-Q., Zhou, H.-L., Lin, R.-B.,
size control in hybrid porous materials for and Chen, X.-M. (2014). Single-crystal X-ray
acetylene capture from ethylene. Science 353, 18. Tan, K., Zuluaga, S., Fuentes, E., Mattson, E.C., diffraction studies on structural transformations
141–144. Veyan, J.-F., Wang, H., Li, J., Thonhauser, T., of porous coordination polymers. Chem. Soc.
and Chabal, Y.J. (2016). Trapping gases in Rev. 43, 5789–5814.
10. Cadiau, A., Adil, K., Bhatt, P.M., Belmabkhout, metal-organic frameworks with a selective
Y., and Eddaoudi, M. (2016). A metal-organic surface molecular barrier layer. Nat. Commun. 27. Huang, R.-W., Wei, Y.-S., Dong, X.-Y., Wu,
framework-based splitter for separating 7, 13871. X.-H., Du, C.-X., Zang, S.-Q., and Mak, T.C.W.

1926 Chem 4, 1911–1927, August 9, 2018


(2017). Hypersensitive dual-function surface corrugation by the use of an aromatic analogues through ligand functionalization.
luminescence switching of a silver- hydrocarbon building unit. Angew. Chem. Int. Inorg. Chem. 53, 9254–9259.
chalcogenolate cluster-based metal–organic Ed. 53, 8225–8230.
framework. Nat. Chem. 9, 689–697. 48. Bai, Y., Dou, Y., Xie, L.-H., Rutledge, W., Li,
38. Liu, X., Li, Y., Ban, Y., Peng, Y., Jin, H., Bux, H., J.-R., and Zhou, H.-C. (2016). Zr-based metal-
28. Burtch, N.C., Jasuja, H., and Walton, K.S. Xu, L., Caro, J., and Yang, W. (2013). organic frameworks: design, synthesis,
(2014). Water stability and adsorption in metal– Improvement of hydrothermal stability of structure, and applications. Chem. Soc. Rev.
organic frameworks. Chem. Rev. 114, 10575– zeolitic imidazolate frameworks. Chem. 45, 2327–2367.
10612. Commun. 49, 9140–9142.
49. Sun, Q., He, H., Gao, W.-Y., Aguila, B., Wojtas,
29. Zhang, W.-X., Liao, P.-Q., Lin, R.-B., Wei, Y.-S., 39. He, C.-T., Jiang, L., Ye, Z.-M., Krishna, R., L., Dai, Z., Li, J., Chen, Y.-S., Xiao, F.-S., and Ma,
Zeng, M.-H., and Chen, X.-M. (2015). Metal Zhong, Z.-S., Liao, P.-Q., Xu, J., Ouyang, G., S. (2016). Imparting amphiphobicity on single-
cluster-based functional porous coordination Zhang, J.-P., and Chen, X.-M. (2015). crystalline porous materials. Nat. Commun. 7,
polymers. Coord. Chem. Rev. 293, 263–278. Exceptional hydrophobicity of a large-pore 13300.
metal-organic zeolite. J. Am. Chem. Soc. 137,
30. Ferey, G., Mellot-Draznieks, C., Serre, C., 7217–7223. 50. Roy, S., Suresh, V.M., and Maji, T.K. (2016). Self-
Millange, F., Dutour, J., Surble, S., and cleaning MOF: realization of extreme water
Margiolaki, I. (2005). A chromium 40. Jayaramulu, K., Datta, K.K.R., Rösler, C., Petr, repellence in coordination driven self-
terephthalate-based solid with unusually large M., Otyepka, M., Zboril, R., and Fischer, R.A. assembled nanostructures. Chem. Sci. 7, 2251–
pore volumes and surface area. Science 309, (2016). Biomimetic superhydrophobic/ 2256.
2040–2042. superoleophilic highly fluorinated graphene
oxide and ZIF-8 composites for oil–water 51. Baertsch, C.D., Funke, H.H., Falconer, J.L., and
31. Huang, X.C., Lin, Y.Y., Zhang, J.P., and Chen, Noble, R.D. (1996). Permeation of aromatic
separation. Angew. Chem. Int. Ed. 55, 1178–
X.M. (2006). Ligand-directed strategy for hydrocarbon vapors through silicalite-zeolite
1182.
zeolite-type metal-organic frameworks: zinc(II) membranes. J. Phys. Chem. 100, 7676–7679.
imidazolates with unusual zeolitic topologies. 41. Jhung, S.H., Lee, J.H., Yoon, J.W., Serre, C.,
Angew. Chem. Int. Ed. 45, 1557–1559. Férey, G., and Chang, J.S. (2007). Microwave
52. Kitagawa, S., Higuchi, M., Kajiwara, T.,
Higashimura, H., Mochizuki, M., Nagashima, K.,
32. Park, K.S., Ni, Z., Cote, A.P., Choi, J.Y., Huang, synthesis of chromium terephthalate MIL-101
and Kiyonaga, T. (2014). Method for producing
R.D., Uribe-Romo, F.J., Chae, H.K., O’Keeffe, and its benzene sorption ability. Adv. Mater.
hydrocarbon, cracking catalyst, porous metal
M., and Yaghi, O.M. (2006). Exceptional 19, 121–124.
complex, method for activating catalyst, and
chemical and thermal stability of zeolitic method for measuring acid strength. JP patent
42. Zhang, K., Lively, R.P., Dose, M.E., Brown, A.J.,
imidazolate frameworks. Proc. Natl. Acad. Sci. application JP2014043435A, filed March 13,
Zhang, C., Chung, J., Nair, S., Koros, W.J., and
USA 103, 10186–10191. 2014.
Chance, R.R. (2013). Alcohol and water
33. Kresge, C.T., Leonowicz, M.E., Roth, W.J., adsorption in zeolitic imidazolate frameworks. 53. Cravillon, J., Münzer, S., Lohmeier, S.-J.,
Vartuli, J.C., and Beck, J.S. (1992). Ordered Chem. Commun. 49, 3245–3247. Feldhoff, A., Huber, K., and Wiebcke, M. (2009).
mesoporous molecular sieves synthesized by a Rapid room-temperature synthesis and
liquid-crystal template mechanism. Nature 43. Ras, M.R., Borrull, F., and Marcé, R.M. (2009).
characterization of nanocrystals of a
359, 710–712. Sampling and preconcentration techniques for
prototypical zeolitic imidazolate framework.
determination of volatile organic compounds
Chem. Mater. 21, 1410–1412.
34. Ben, T., Ren, H., Ma, S., Cao, D., Lan, J., Jing, X., in air samples. Trac-trend Anal. Chem. 28,
Wang, W., Xu, J., Deng, F., Simmons, J.M., 347–361. 54. Kayal, S., Sun, B., and Chakraborty, A. (2015).
et al. (2009). Targeted synthesis of a porous Study of metal-organic framework MIL-101(Cr)
aromatic framework with high stability and 44. Bon, V., Senkovskyy, V., Senkovska, I., and for natural gas (methane) storage and compare
exceptionally high surface area. Angew. Chem. Kaskel, S. (2012). Zr(IV) and Hf(IV) based metal– with other MOFs (metal-organic frameworks).
Int. Ed. 48, 9457–9460. organic frameworks with reo-topology. Chem. Energy 91, 772–781.
Commun. 48, 8407–8409.
35. Spek, A.L. (2009). Structure validation in 55. Oxford Diffraction. (2010). CrysAlis Pro
chemical crystallography. Acta Crystallogr. D 45. Bon, V., Senkovska, I., Baburin, I.A., and Kaskel, Software, Ver. 1.171.34 (Oxford Diffraction Ltd).
Biol. Crystallogr. 65, 148–155. S. (2013). Zr-and Hf-based metal–organic
frameworks: tracking down the polymorphism. 56. Sheldrick, G.M. (2008). A short history of
36. Valenzano, L., Civalleri, B., Chavan, S., Bordiga, Cryst. Growth Des. 13, 1231–1237. SHELX. Acta Crystallogr. A. 64, 112–122.
S., Nilsen, M.H., Jakobsen, S., Lillerud, K.P., and
Lamberti, C. (2011). Disclosing the complex 46. Furukawa, H., Gándara, F., Zhang, Y.-B., Jiang, 57. Toth, J. (1995). Uniform interpretation of gas-
structure of UiO-66 metal organic framework: a J., Queen, W.L., Hudson, M.R., and Yaghi, O.M. solid adsorption. Adv. Colloid Interf. Sci. 55,
synergic combination of experiment and (2014). Water adsorption in porous metal– 1–239.
theory. Chem. Mater. 23, 1700–1718. organic frameworks and related materials.
J. Am. Chem. Soc. 136, 4369–4381. 58. Pan, H.H., Ritter, J.A., and Balbuena, P.B.
37. Rao, K.P., Higuchi, M., Sumida, K., Furukawa, (1998). Examination of the approximations
S., Duan, J., and Kitagawa, S. (2014). Design of 47. Wang, B., Huang, H., Lv, X.-L., Xie, Y., Li, M., used in determining the isosteric heat of
superhydrophobic porous coordination and Li, J.-R. (2014). Tuning CO2 selective adsorption from the Clausius-Clapeyron
polymers through the introduction of external adsorption over N2 and CH4 in UiO-67 equation. Langmuir 14, 6323–6327.

Chem 4, 1911–1927, August 9, 2018 1927

You might also like