Tải bản đầy đủ (.pdf) (12 trang)

Gas sorption properties and kinetics of porous bismuth-based metal-organic frameworks and the selective CO2 and SF6 sorption on a new bismuth trimesate-based structure UU-200

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (6.39 MB, 12 trang )

Microporous and Mesoporous Materials 329 (2022) 111548

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials
journal homepage: www.elsevier.com/locate/micromeso

Gas sorption properties and kinetics of porous bismuth-based metal-organic
frameworks and the selective CO2 and SF6 sorption on a new bismuth
trimesate-based structure UU-200
Michelle Åhl´en a, Elina Kapaca b, Daniel Hedbom a, Tom Willhammar b, Maria Strømme a,
Ocean Cheung a, *
a

Division of Nanotechnology and Functional Materials, Department of Materials Science and Engineering, Uppsala University, ngstră
om Laboratory, Uppsala, SE, 751
03, Box 35, Sweden
b
Department of Materials and Environmental Chemistry, Stockholm University, Stockholm, SE, 106 91, Sweden

A R T I C L E I N F O

A B S T R A C T

Keywords:
Bismuth
Metal-organic frameworks
Porosity
Greenhouse gas capture
Adsorption kinetics


Bismuth-based metal-organic frameworks (Bi-MOFs) such as bismuth subgallate are important for applications
ranging from medicine to gas separation and catalysis. Due to the porous nature of such Bi-MOFs, it would be
valuable to understand their gas sorption and separation properties. Here, we present the gas sorption properties
of three microporous Bi-MOFs, namely, CAU-17, CAU-33, and SU-101, along with a new trimesate-based
structure, UU-200. We perform a detailed analysis of the sorption properties and kinetics of these Bi-MOFs.
UU-200 shows good uptake capacities for CO2 (45.81 cm3 g− 1 STP) and SF6 (24.69 cm3 g− 1 STP) with CO2/
N2 and SF6/N2 selectivities over 35 and 44, respectively at 293 K, 100 kPa. The structure of UU-200 is inves­
tigated using continuous rotation electron diffraction and is found to be a 3D porous framework containing pores
with a diameter of 3.4–3.5 Å. Bi-MOFs as a group of relatively under-investigated types of MOFs have interesting
sorption properties that render them promising for greenhouse gas adsorbents with good gas uptake capacities
and high selectivities.

1. Introduction
The emission of greenhouse gases (GHGs), and in particular carbon
dioxide (CO2), has become an ever-increasing concern in today’s society
as global warming and climate change-related issues become more ur­
gent [1–3]. Efforts have been made in the last couple of decades to
reduce the anthropogenic emission of CO2 through investments into
renewable energy sources, efficiency improvements, and low-carbon
fuels, to name a few [4,5]. Carbon capture and storage (CCS) technol­
ogies have garnered significant attention in the last couple of decades as
a potential low-cost and facile alternative for CO2 sequestration through
the use of solid microporous sorbents such as zeolites [6,7], porous
carbons [8,9], and metal-organic frameworks (MOFs) [10,11]. CO2
capture and separation through the use of liquid amine solutions (also
known as amine scrubbing) has been utilized in industrial plants since
the 1930s [12]. The corrosivity and volatility along with high-energy
requirements and cost for recycling the amine-based solutions impose
certain limitations on this CCS technology. However, contrary to the


traditional amine-based absorption methods, solid sorbents present
potential advantages such as reduced regeneration energy requirements,
improved ease of handling, high adsorption capacities, and good sepa­
ration performances, to name a few [13]. In particular MOFs, a diverse
and relatively new class of functional porous materials, have attracted
attention as promising sorbents for greenhouse gas capture and sepa­
ration [14]. The structural diversity of MOFs arises due to the wide
range of metal cations (or metal clusters), also known as secondary
building units (SBUs), and organic linkers that can be combined to form
2D and 3D frameworks of various topologies. Enabling the formation of
framework materials with tunable pore sizes and shapes, surface func­
tionalities, and physical properties [2,10], making these materials
interesting for various applications beyond CCS technologies, such as in
drug delivery, catalysis, energy conversion, gas sensing, and
luminescence-based sensing [15–20]. Many MOF structures containing
metals from the s-, p-, d-, and f-blocks have been synthesized over the
years [21,22]; however, bismuth-based MOFs have remained relatively
scarce. Organometallic complexes such as bismuth subgallate, an active

* Corresponding author.
E-mail address: (O. Cheung).
/>Received 18 August 2021; Received in revised form 8 October 2021; Accepted 2 November 2021
Available online 9 November 2021
1387-1811/© 2021 The Authors. Published by Elsevier Inc. This is an open access article under the CC BY license ( />

M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

pharmaceutical ingredient used to treat mild gastrointestinal ailments,

have been shown to have noticeable CO2 and N2 porosities [23], how­
ever, only a handful of bismuth complexes have been observed to be
highly microporous [24,25]. The emergence of the first permanently
porous Bi-MOF, CAU-7, composed of 1,3,5-benzenetrisbenzoate with a
recorded specific surface area of 1150 m2 g 1, was introduced by Norư
bert Stocks group at Christian-Albrechts-Universită
at (= CAU) in 2012
[26]. Porous bismuth-based MOFs composed of various other organic
linkers have followed during the last decade, such as NOTT-220 (3,3′ ,5,
5′ -tetracarboxylate-based)
[27],
CAU-35
(triazine-2,4,6-triyl-­
tribenzoate-based) [28], CAU-17 (1,3,5-benzenetricarboxylate-based)
[29], CAU-33 (1,2,4,5-tetrakis-(4-carboxyphenyl)benzene-based) [30],
Bi-NU-901 (1,3,5,8-(p-benzoate)pyrene-based) [31] and SU-101 (ella­
gate-based) [32], to name a few. The trimesate-based MOF CAU-17 has
been shown to possess an intricate topological structure [33] and
Bi-MOFs with various structures have been synthesized using 1,3,5-ben­
zenetricarboxylic acid as the organic linker [34–37]. As such, due to
their structural versatility trimesate-based Bi-MOFs could prove to be
interesting sorbents for greenhouse capture applications.
Herein, we present a detailed analysis of the gas sorption properties
of a new Bi-MOF, UU-200 (UU = Uppsala University), synthesized from
1,3,5-benzenetricarboxylic acid and Bi(NO3)3⸱5H2O, and three micro­
porous bismuth-based MOFs; CAU-17, CAU-33, and SU-101. The struc­
ture of UU-200 was studied using a 3-dimensional electron diffraction
(3DED/MicroED) technique along with powder X-ray diffraction
(PXRD). The porosities and greenhouse gas capture properties of the BiMOFs were investigated using a volumetric equilibrium-based sorption
method and the CO2 adsorption kinetics were studied using a

gravimetric-based technique. The rate-limiting mechanisms governing
the CO2 adsorption process were investigated using the obtained
gravimetric adsorption profiles and estimated CO2 diffusivities were
calculated.

Teflon-lined stainless-steel autoclave and heated at 120 ◦ C for 12 h. The
obtained white product was collected from the cooled autoclave by
centrifugation at 3800 rpm for 20 min, washed with 40 ml MeOH three
times, and dried in a ventilated oven at 70 ◦ C overnight.
2.2.3. Synthesis of CAU-33
CAU-33 was prepared according to a previously reported procedure
[30]. Briefly, Bi(NO3)3⸱5H2O (174 mg, 358 μmol) and H4TCPB (100 mg,
179 μmol) were dissolved in a mixture of 4.5 ml DMF and 0.5 ml toluene
and transferred to a 25 ml Teflon-lined stainless steel autoclave. The
autoclave was heated at 120 ◦ C for 12 h and thereafter left to cool to
ambient temperatures naturally. The obtained product was
solvent-exchanged in a 1:1 mixture of MeOH and DMF at 100 ◦ C for 10
min, washed with MeOH three times, and dried in a ventilated oven at
70 ◦ C for 40 min.
2.2.4. Synthesis of SU-101
SU-101 was synthesized according to procedures reported by
Svensson Grape et al. [32]. Briefly, BiAc3 (380 mg, 1.0 mmol) and ellagic
acid dihydrate (150 mg, 0.5 mmol) were dispersed in a mixture of 1.8 ml
concentrated acetic acid and 28.3 ml deionized water. The dispersion
was left stirring at room temperature for 48 h, after which the product
was separated by centrifugation at 3800 rpm for 20 min, washed with
deionized water once and EtOH twice, and finally dried in a ventilated
oven at 70 ◦ C overnight.
2.3. Characterization
Powder X-ray diffraction (PXRD) diffractograms were recorded on a

Bruker D8 Advance Powder diffractometer (Bruker, Bremen, Germany)
operated at 40 kV and 40 mA, using Cu Kα radiation (λ = 1.5418 Å), a
step-size of 0.02◦ and a time-per-step of 0.3 s. PXRD data for Rietveld
refinement of UU-200 were collected using a Panalytical X’Pert alpha1
powder X-ray diffractometer equipped with Johansson Ge mono­
chromator producing Cu-Kα1 radiation (λ=1.540598 Å). Scanning
electron microscopy (SEM) images were obtained on a Zeiss Merlin Field
Emission Scanning Electron Microscope (Oberkochen, Germany) using
an acceleration voltage of 2.5 kV and a probe current of 80 pA. All
samples were sputter-coated with a thin layer of Pd/Au prior to imaging.
Fourier transform infrared (FTIR) spectra were recorded on a Bruker
Tensor 27 (Bruker, Bremen Germany) using a platinum-attenuated total
reflectance (ATR) accessory. Gas sorption experiments were carried out
on a Micromeritics ASAP 2020 surface area analyzer (Norcross, GA,
USA) and all samples were degassed at 423 K for 3 h in dynamic vacuum
(1 × 10− 4 Pa) before analysis. Brunauer-Emette-Teller (BET) and
Langmuir specific surface areas were calculated from N2 isotherms
recorded at 77 K at 4.5–16 kPa and p/p0 = 0.05–0.15, respectively. Pore
size distributions were calculated using the Density Functional Theory
(DFT) function in the Micromerities MicroActive software using the N2
isotherms, the slit pore model for N2 was used for these calculations.
Total pore volumes were determined from N2 and H2O isotherms
recorded at 77 K and 293 K, respectively, using a single point from the
adsorption branch at 0.98 and 0.93 p/p0. These p/p0-values were chosen
(instead of 0.99 in both cases) to avoid possible overdosing of the
samples due to experimental error, which would cause the condensation
of the adsorbate gas and hence overestimating the pore volume. Sorp­
tion isotherms of CH4, CO2, N2, and SF6 between 273 and 303 K were
also obtained using the Micromeritics ASAP 2020 surface area analyzer
(Norcross, GA, USA) but with an insulating water bath containing either

water or a water-ice slurry. The gas selectivities were calculated for
theoretical gas mixtures containing CH4/N2 (50:50), CO2/CH4 (50:50),
CO2/N2 (85:15), and SF6/N2 (10:90) using s = (qgas1/qgas2)/(Pgas1/Pgas2)
and the Ideal Adsorption Solution Theory (IAST). Single-component
isotherms of CH4, CO2, N2, and SF6 recorded at 293 K were used for
the IAST calculations (see Section 7 in Supporting Information for more
details) and all isotherms were fitted with the single-site Langmuir

2. Experimental
2.1. Materials
Bismuth(III) nitrate pentahydrate (Bi(NO3)3⸱5H2O), 1,3,5-benzene­
tricarboxylic acid (Trimesic acid, H3BTC), 1,2,4,5-tetrakis-(4-carboxy­
phenyl)benzene (H4TCPB), and Acetic acid ≥99% were purchased
from Sigma-Aldrich, USA. N,N-Dimethylformamide (DMF), Methanol
(MeOH), Ethanol (EtOH), Toluene, Ellagic acid dihydrate, and Bismuth
(III) acetate (BiAc3) were purchased from VWR International AB,
Sweden.
All chemicals were used as obtained without further purification.
2.2. Synthesis
2.2.1. Synthesis of UU-200
In a typical synthesis, Bi(NO3)3⸱5H2O (454 mg, 936.8 μmol) and
H3BTC (957 mg, 4.6 mmol) were dissolved in 10 ml and 15 ml of DMF,
respectively. The two solutions were thereafter mixed, by the addition of
the metal salt solution to the H3BTC solution, after which the mixture
was transferred to a 50 ml Teflon-lined stainless-steel autoclave. The
autoclave was heated at 140 ◦ C for 72 h and left to cool to room tem­
perature. The obtained product was collected by centrifugation at 3800
rpm, washed once with 40 ml DMF and solvent-exchanged in a 50 ml
solution of MeOH and DMF (1:1 v/v) at 100 ◦ C for 20 min. Lastly, the
product was further washed with 40 ml MeOH twice and finally dried

overnight in a ventilated oven at 70 ◦ C.
2.2.2. Synthesis of CAU-17
CAU-17 was prepared with an adaption of a previously reported
procedure [29]. Briefly, a solid mixture of Bi(NO3)3⸱5H2O (50 mg, 103.1
μmol) and H3BTC (250 mg, 1.2 mmol) was dissolved in 50 ml MeOH.
The clear and homogenous mixture was transferred to a 50 ml
2


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

model, dual-site Langmuir model, or Toth model. Isosteric enthalpies of
adsorption (-ΔHads) were calculated using the Clausius-Clapeyron
equation on CO2 and SF6 adsorption isotherms recorded at 273, 283,
293, and 303 K and fitted using the dual-site Langmuir model or Toth
model. Gravimetric adsorption profiles were recorded using a Mettler
Toledo TGA/DSC 3+ (Schwerzenbach, Switzerland) using N2 as purge
gas and CO2 as sorbate. All experiments were carried out on 2.5–7.5 mg
of material degassed at 423 K for 30 min in an N2 atmosphere (50 ml
min− 1 flow rate) prior to the gas being switched to CO2, which occurred
at 303 K and proceeded for 20 min (50 ml min− 1 flow-rate). The ob­
tained CO2 adsorption profiles were further studied using three diffusion
models and estimated CO2 diffusivities were calculated (see Section 9 in
Supporting Information for details).

performed by direct methods using software SIR2014 [40] and the
structure refinement was performed using SHELXL-97 [41].
3. Results and discussion

3.1. Structure of UU-200
While the structure of CAU-17, CAU-33, and SU-101 were reported in
recent literature [29,30,32,33], the structure of UU-200 has not been
previously reported. The cRED data from UU-200 could be indexed
using an orthorhombic unit cell: a = 22.51 Å, b = 27.53 Å, and c = 10.42
Å, and space group Pnnm (No. 58), see Fig. 1. A Pawley fit against
in-house PXRD data was performed and confirmed the space group and
unit cell parameters a = 21.6381 Å, b = 27.9108 Å, and c = 9.8961 Å
(Fig. S1). The structure of UU-200 was determined from cRED data of
88% completeness with a resolution up to 1.2 Å and Rint = 0.35, using
unit cell parameters obtained from the Pawley fit. The structure solution
resulted in three bismuth, eight oxygen, and 27 carbon atoms in the
asymmetric unit. Eight additional oxygen atoms were located from the
difference electrostatic potential map during refinement to complete the
structure. The structure refinement converged with an R1-value of 0.33
and GooF of 2.23, see Table S1 for complete statistics.
The structure of UU-200 was also confirmed by refinement against
PXRD data. Rietveld refinement converged with Rwp = 0.205, further
details regarding the Rietveld refinement can be found in SI (Table S2
and Fig. S4). UU-200 crystallizes in space group Pnnm (No. 58) and
exhibits a 3D structure containing three Bi3+ ions and four 1,3,5-benze­
netricarboxylate (BTC3− ) anions in the asymmetric unit. The Bi13+ and
Bi33+ ions are coordinated with 10 oxygen atoms from six different
BTC3- anions, and Bi23+ is coordinated with nine oxygen atoms from five

2.4. Structure determination
Continuous rotation electron diffraction (cRED) was used to deter­
mine the structure of UU-200. The sample was prepared by crushing the
powder in agate mortar and dispersing it in absolute ethanol. A droplet
of the suspension was transferred to a copper grid covered with a holey

carbon film. cRED data were collected using a JEOL JEM-2100 trans­
mission electron microscope (TEM) (Akashima, Japan). The TEM was
operated at 200 kV and the sample was mounted on a cryo-transfer to­
mography sample holder (Gatan 914) and cooled to 98 K using liquid N2
during the data collection. The cRED data were collected by continu­
ously tilting the goniometer and registered using a high-speed hybrid
detection camera (Timepix Quad, ASI) using the software Instamatic
[38]. 3D ED datasets were processed using the software XDS [39]. In­
tegrated reflection intensities from three datasets were merged and used
for structure determination. The structure solution of UU-200 was

Fig. 1. (a) Projection of the 3D reciprocal lattice reconstructed from the cRED data with an inset of the crystal from which data were collected. 2D slices from the 3D
data including the (b) 0kl, (c) hk0, and (d) h0l families of reflections.
3


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

different BTC3− anions (Fig. 2d). The structure of UU-200 contains
bismuth atoms that are connected to form Bi2 secondary building units
(SBUs) (Fig. 2d). They are connected via three (for Bi2 and Bi3) and four
(for Bi1) oxygen atoms from different BTC3− anions. The distance be­
tween Bi2 and Bi3 is 4.04 Å and the distance between two Bi1 atoms in a
Bi2 SBU is 3.99 Å. Bi2 SBUs have previously been reported for two Bi3+MOF structures, namely, Bi-BTC [34,35] and SU-100 [42]. Due to the
difference in coordination environment, they all exhibit different
structures. UU-200 exhibits a 3D pore system limited by small windows.
Diffusion along the c-axis will be limited by a triangular pore with a size
of 3.49 Å, and diffusion along the a- and b-axes will both be limited by a

3.42 Å window, see Fig. S3 (all sizes are given after subtraction of the
van der Waals radii, 1.35 Å for oxygen and 1.70 Å for carbon). Which
was found to be smaller than the crystallographic pore apertures ob­
tained from CAU-17 (9.6 Å, 3.6 Å, and 3.4 Å) [33], CAU-33 (9.5 × 4.6 Å
and 4.4 × 4.1 Å) [30], and SU-101 (6 - 7 Å) [32].
Each of the four symmetry-independent BTC3− anions are in a
different chemical environment, see Fig. S2. One of the BTC3− anions
coordinates with all three carboxylate groups in a bidentate chelating
mode to a Bi3+, a second BTC3− anion coordinates with two of the
carboxylate groups in a chelating mode to each one Bi3+ ion while the
remaining carboxylate chelate to one Bi3+ and forms an additional bond
to one additional Bi3+. The remaining two BTC3− anions chelate with
one carboxylate group to a Bi3+ while the last two carboxylate groups
form three bonds each to two different Bi3+-ions. IR spectra of UU-200
(Fig. S9) also show a clear interaction between the Bi3+ ions and the
BTC3− linker, as indicated by a slight blue-shift of the carboxylate group
of in the linker from approx. 1694 cm− 1 to slightly higher wavenumbers.
The structures of UU-200 and CAU-17 are both built from Bi3+ ions
connected by BTC3− anions. Their crystal structures are however
different. In CAU-17, each Bi3+ (there are nine symmetry independent
Bi3+) is coordinated by nine oxygen atoms - eight of these oxygen atoms
belong to carboxylate groups of the BTC3− ions and the ninth is a
coordinating water molecule. This is different from the structure of UU200 as described above. The crystal structure of UU-200 is also different
from that of the intermediate [Bi(HBTC) (NO3) (MeOH)]MeOH phase
ăppen et al. [29] which contains coordinating NO3− ions.
identified by Ko

3.2. Porosity of Bi-MOFs by N2 and H2O sorption
Nitrogen sorption isotherms and the calculated specific surface areas
(SSAs) (Fig. 3 and Table 1) show that the four studied Bi-MOFs were

porous towards N2 at 77 K. According to the structural study, the pores
of UU-200 were found to be comparable to the kinetic diameter of N2. In
fact, very slow N2 diffusion was observed in UU-200 while recording the
N2 sorption isotherm (with an excess equilibration time required at low
pressures). Indicating the presence of very narrow pores on UU-200.
Pores with effective diameters close to the kinetic diameter of N2 (or
other adsorbates of choice) would effectively restrict the diffusion of N2
into the pores. This could result in an extensive amount of time needed
for an equilibrium adsorption point to be recorded during the gas
adsorption experiments, or severely restrict N2 diffusion (i.e. in zeolite

Fig. 3. Nitrogen sorption isotherms of UU-200, CAU-17, CAU-33, and SU-101
recorded at 77 K. The adsorption and desorption branches of the isotherms
are indicated by filled and hollow symbols, respectively.

Fig. 2. The structure of UU-200 viewed along the (a) a-, (b) b- and (c) c-axes, respectively, d) the two unique Bi2 SBUs of the UU-200 structure. Bismuth is shown in
purple, oxygen in red, and carbon in grey. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
4


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

Table 1
Langmuir and BET specific surface areas and pore volumes of UU-200, CAU-17,
CAU-33, and SU-101. The values for UU-200 (in brackets) are likely to be
underestimated due to the pore-size limited adsorption of N2.
Sample


SSALangmuira
(m2 g− 1)

SSABETb
(m2 g− 1)

Vmicroc
(cm3 g− 1)

Vtotd
(cm3
g− 1)

Vtot, H2Oe
(cm3 g− 1)

UU200
CAU17
CAU33
SU-101

(141)

(115)



(0.07)

0.13


610

496

0.22

0.22

0.20

459

373

0.15

0.19

0.16

426

344

0.13

0.22
(0.17)f


0.13

a
Langmuir specific surface areas (SSALangmuir) were calculated using the
Langmuir equation between 4.5 and 16 kPa.
b
BET specific surface areas (SSABET) were calculated using the BrunauerEmmett-Teller (BET) equation between 0.05–0.15 p/p0.
c
Micropore volumes (Vmicro) were estimated using the t-plot method.
d
Total pore volumes (Vtot) were determined using a single-point from the
adsorption branch of the isotherm at p/p0 = 0.98.
e
Total pore volumes (Vtot) were determined at 293 K using a single-point from
the adsorption branch of the H2O isotherm at p/p0 ≥ 0.93.
f
Vtot determined using a single-point from the adsorption branch of the
isotherm at p/p0 = 0.80.

Fig. 4. Water sorption isotherms recorded at 293 K and 100 kPa for UU-200,
CAU-17, CAU-33, and SU-101. The adsorption and desorption branches of the
isotherms are indicated by filled and hollow symbols, respectively.

at pressures below 20 kPa, followed by adsorption in the larger hexag­
onal pores above 20 kPa, in the framework [29]. Similarly, the hysteresis
that is observed for CAU-33 may be connected to a difference in H2O
desorption rate from the larger (9.4 × 4.6 Å) and smaller (4.4 × 4.1 Å)
1D channels in the material [30]. The Langmuir-shaped isotherm of
UU-200 points toward an enhanced H2O affinity to the materials as
compared to CAU-33 and SU-101. This enhanced affinity is assumed to

be due to the presence of more suitable sized pores in UU-200 and
CAU-17, and not due to a significant difference in framework hydro­
philicity. As the organic linker in CAU-33 can be assumed to have a
similar hydrophobic character to the BTC-linker, while the
ellagate-linker in SU-101 may show slightly higher hydrophilic prop­
erties due to the lactone ring on the ligand [32]. The calculated pore
volume of SU-101 was found to be significantly lower as compared to
those estimated made from N2 sorption. This was discrepancy may be
attributed to an intraparticle condensation of N2 at relative pressures
above p/p0 = 0.80, or possible mesoporosity that arises from structural
defects. Determination of the total pore volume at a p/p0 of 0.80 resulted
in a pore volume that was comparable to that obtained from H2O
sorption (Table 1). This increased affinity is likely related to the pore
size of UU-200. A comparison between the calculated pore volumes
determined from the N2 isotherms at 77 K and the H2O isotherms
recorded at 293 K (Table 1) shows very different porosities for the MOFs.
While UU-200 was found to have low porosity as determined by N2, the
obtained porosities from the H2O sorption isotherms show UU-200 to be
comparable to the other three Bi-MOFs (given that the hydrophilicity of
the frameworks can be assumed to be somewhat comparable).

3A [7]), in both cases, the specific surface area and pore volume of the
material would be underestimated. Similar to what was observed
recently on some mixed-linker ZIF-7-8s [43]. As a result, N2 sorption
points at very low relative pressures were omitted during the recording
of the isotherm. The values listed in Table 1 for UU-200 are therefore
expected to be underestimates of the true values. Micropore analysis
(which requires data points at low relative pressures) by N2 sorption
using the same settings as those for the other Bi-MOFs was therefore not
performed on UU-200. On the other hand, CO2 sorption isotherms

(discussed in detail later) of UU-200 suggested that the structure has
comparable porosity to CAU-17, CAU-33, and SU-101.
The DFT-PSD of CAU-17, CAU-33, and SU-101 (Fig. S12) were found
to be in good agreement with previously published data [29,30,32] and
all three Bi-MOFs were found to be microporous (pores of <2 nm in
diameter [44]). Discrepancies between the estimated size of the pore
apertures in this study as compared to previous literature can most likely
be attributed to a difference in crystallinity (see Fig. S5) and different
degrees of pore-evacuation prior to the N2 sorption measurements. The
pore size calculation on UU-200 was not performed, as previously
mentioned, due to the slow diffusion of N2. Detailed discussion related to
the limit of N2 sorption on narrow pores could be found in our previous
work [43].
As N2 could not be used as an adsorbate to probe the porosity of UU200, water (H2O) sorption isotherms were recorded for all Bi-MOFs at
293 K. According to the isotherms shown in Fig. 4, H2O as a small
adsorbate (kinetic diameter of 2.8 Å [45] compared to 3.6 Å for N2 [46])
was able to freely adsorb on all of the Bi-MOFs, including UU-200. The
porosity of UU-200 could be qualitatively demonstrated using H2O as
adsorbate and the respective pore volumes of the Bi-MOFs are presented
in Table 1. However, as H2O molecules would interact with each other,
and possibly with the Bi-MOF frameworks in the free and adsorbed state,
calculations of the SSAs of the Bi-MOFs using the H2O sorption data were
not performed due to the complexity of using water as an adsorbate
(water molecules are expected to interact with each other and therefore,
deviate from the assumptions made by the BET and Langmuir models).
The shape of the isotherms were found to differ significantly between
the different MOFs. A Langmuir-shaped isotherm was observed for
UU-200 and CAU-17. The step in the adsorption branch of the isotherm
for CAU-17 has previously been found to correspond to a sequential H2O
adsorption in the smaller triangular and rectangular pores taking place


3.3. Greenhouse gas adsorption and separation on Bi-MOFs
The Bi-MOFs synthesized in this study were examined for their
ability to separate and adsorb greenhouse gases. Equilibrium gas sorp­
tion isotherms were recorded at 293 K on the synthesized Bi-MOFs using
CH4, CO2, N2, and SF6 as adsorbed gases (Fig. 5). The gas uptake ca­
pacities at 100 kPa are listed in Table 2. Due to the restricted N2 diffu­
sion observed on UU-200 at 77 K, the N2 adsorption isotherms at 293 K
were further investigated in order to evaluate the pore size effect. The N2
sorption capacity of UU-200 at 100 kPa (3.59 cm3 g− 1 STP) was found to
be comparable to CAU-17 (4.78 cm3 g− 1 STP) and SU-101 (2.36 cm3 g− 1
STP), despite the noticeably higher N2-porosity on the latter two MOFs.
This indicated that the limited N2 access observed on UU-200 at 77 K,
which was attributed to the presence of narrow pores, could in fact also
5


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

Fig. 5. CH4, CO2, N2, and SF6 sorption isotherms recorded at 293 K and 100 kPa for (a) UU-200, (b) CAU-17, (c) CAU-33, and (d) SU-101. The adsorption and
desorption branches of the isotherms are indicated by filled and hollow symbols, respectively.
Table 2
CH4, CO2, N2, and SF6 uptake capacities and gas selectivities at 293 K and 100 kPa for UU-200, CAU-17, CAU-33, and SU-101. Working capacity is calculated for SF6 between 10 and 100 kPa and CO2 - between 15 and 100 kPa).
Sample

UU200
CAU17
CAU33

SU-101

Working capacity

IAST selectivity

Henry’s law selectivity

CH4 (cm3
g− 1 STP)

CO2 (cm3
g− 1 STP)

N2 (cm3
g− 1 STP)

SF6 (cm3
g− 1 STP)

CO2 (cm3
g− 1 STP)

SF6 (cm3
g− 1 STP)

CO2/CH4
(50:50)

CO2/N2

(15:85)

SF6/N2
(10:90)

CO2/
CH4

CO2/
N2

SF6/
N2

11.19

45.81

3.59

24.69

8.28

26.66

5.39

34.52


44.81

10.99

29.52

22.25

20.75

52.74

4.78

32.47

15.01

33.15

1.60

26.13

35.58

4.79

21.47


18.33

17.69

43.95

4.93

34.84

26.21

23.30

3.11

27.35



5.16

23.82

27.97

7.41

47.82


2.36

17.91

6.50

32.26

8.34

36.29

39.87

10.84

32.68

21.59

be related to a mild thermally responsive structural change in the
framework. Negative thermal expansions (NTE), i.e. a reduction in the
crystallographic unit cell with increasing temperature, has been
observed in MOFs such MOF-5 [47], IRMOFs [48], Cu3BTC2 [49], and
UiO-66(Hf) [50], while positive thermal expansion (PTE) effects, i.e. an
increase in the crystallographic unit cell with increasing temperature,
has been detected to a lesser degree in framework materials [51].
However, both NTE and PSE have been observed in DMF-solvated
DUT-49(Cu) [51,52], where a ~6.4% reduction in the unit cell vol­
ume was observed upon cooling the material from 298 K to 150 K. This

phenomenon was attributed to the solidification of the DMF molecules
in the pores of the structure resulting in a contraction of the pore vol­
ume. Similar observations were made in two DMF-solvated Mn- and
Cd-based MOFs [53,54], which both showed low N2-porosities at 77 K
and both PTE and NTE properties, resulting in a slight increase in unit
cell volume between 208 and 215 K (corresponding to the melting point
of DMF at 212 K). Although no evidence of framework flexibility was
observed during the sorption of various gases at 293 K (e.g. CH4, CO2,

N2, or SF6, see Fig. 5 and Table 2). The possibility that UU-200 may
undergo a significant enough thermal expansion between 77 K and 293
K for the structure to become N2-porous at ambient temperatures could
explain the conflicting results that were obtained.
The CO2 uptakes on all Bi-MOFs were found to be notable with CAU17 showing the highest uptake of 57.74 cm3 g− 1 STP at 100 kPa (293 K),
followed by SU-101 (47.82 cm3 g− 1 STP), UU-200 (45.81 cm3 g− 1 STP),
and CAU-33 (43.95 cm3 g− 1 STP). The CO2 uptake on UU-200 as well as
the other three Bi-MOFs was found to be comparable to UiO-66(Zr)
(49.31 cm3 g− 1 STP at 273 K and 100 kPa), SIFSIX-2-Cu (41.24 cm3
g− 1 STP at 298 K and 100 kPa), USTC-253 (47.74 cm3 g− 1 STP at 298 K
and 100 kPa), MIL-125 (48.86 cm3 g− 1 STP at 298 K and 100 kPa), UPC105 (53.12 cm3 g− 1 STP at 298 K and 100 kPa), and SIFSIX-3-Ni (60.52
cm3 g− 1 STP at 298 K and 100 kPa) [55]. The CO2 adsorption capacities
were however higher than the reported CO2 uptakes on MOF-177
(17.26 cm3 g− 1 STP at 298 K and 100 kPa), SNU-70 (17.93 cm3 g− 1
STP at 298 K and 100 kPa), MIL-101(Cr) (27.35 cm3 g− 1 STP at 298 K
and 100 kPa), and MIP-202 (12.33 cm3 g− 1 STP at 298 K and 100 kPa)
6


M. Åhl´en et al.


Microporous and Mesoporous Materials 329 (2022) 111548

and lower than HKUST-1 (93.24 cm3 g− 1 STP at 298 K and 100 kPa),
PCN-88 (93.91 cm3 g− 1 STP at 298 K and 100 kPa), Mg-MOF-74 (179.31
cm3 g− 1 STP at 296 K and 100 kPa), Co-MOF-74 (156.90 cm3 g− 1 STP at
296 K and 100 kPa), Ni-MOF-74 (130.00 cm3 g− 1 STP at 296 K and 100
kPa), and ZJNU-44 (116.10 cm3 g− 1 STP at 296 K and 100 kPa)
(Table S3) [55]. A number of observations could be made based on the
CO2 sorption isotherms of the Bi-MOFs, namely: 1) All of the CO2 iso­
therms adopted the Langmuir-shape, implying that the effective pore
size of these Bi-MOFs are close to the kinetic diameters of CO2 for
enhanced sorbent-sorbate interaction (as CO2 does not chemisorb on
Bi-MOFs). 2) The low N2 porosity observed on UU-200 was not reflected
by the CO2 uptake amount at 293 K – the CO2 adsorption capacity on
UU-200 was comparable to other Bi-MOFs with high N2 surface areas (e.
g. CAU-17 and SU-101). This confirmed that the low apparent N2
porosity on UU-200 was due to kinetically restricted N2 diffusion, which
might be due to a narrowing of the pore opening on UU-200 at 77 K. 3)
The two trimesate-based Bi-MOFs; UU-200 and CAU-17, although
chemically similar, had CO2 uptake capacities that were noticeably
different, which is most likely related to the structural differences be­
tween the materials, as was previously mentioned. Low-pressure CO2
uptakes (Fig. S13) at industrially relevant pressures (i.e. 15 kPa) were
found to follow a similar trend to those observed at 100 kPa. The
adsorption capacity of CAU-17 (20.62 cm3 g− 1 STP) and CAU-33 (20.62
cm3 g− 1 STP) were observed to be equivalent while UU-200 (19.05 cm3
g− 1 STP) and SU-101 (16.14 cm3 g− 1 STP) had slightly lower uptakes.
The SF6 uptake on all Bi-MOFs are also presented in Fig. 5 and
Table 2. In all cases, SF6 uptake on Bi-MOFs was higher than the N2
uptake, this is due in part to the higher boiling point of SF6 (223 K) than

N2 (77 K). The high boiling point of SF6 means that SF6 condenses more
readily than N2 on surfaces. the highest SF6 uptake capacity was
observed on CAU-33 (34.84 cm3 g− 1 STP), followed by CAU-17 (32.47
cm3 g− 1 STP), UU-200 (24.69 cm3 g− 1 STP), and SU-101 (17.91 cm3 g− 1
STP). The SF6 uptake on CAU-17 synthesized in this study matched that
ăppen et al. (~31.38 cm3 g 1 STP at 293 K
of CAU-17 synthesized by Ko
and 100 kPa) [29]. The uptake on UU-200 was higher than expected as
the crystallographic pore window of the MOF was significantly smaller
than the kinetic diameter of SF6 (5.5 Å). Thus, it is possible that the SF6
adsorption may be related to a thermal expansion and/or to structural
flexibility of the framework, although future investigations including
possibly the use of in-situ PXRD technique could be interesting in
studying the sorption properties of UU-200. The SF6 uptakes on the
Bi-MOFs were comparable to other microporous sorbents such as MOF-5
(39.00 cm3 g− 1 STP at 298 K and 100 kPa), UiO-66(Zr) (32.50 cm3 g− 1
STP at 298 K and 100 kPa), and Zeolite 13X (39.22 cm3 g− 1 STP at 298 K
and 100 kPa) [56]. On the other hand, other MOFs have been shown to
have higher SF6 uptakes than the Bi-MOFs in this study, including
Mg-MOF-74 (141.21 cm3 g− 1 STP at 298 K and 100 kPa), Co-MOF-74
(116.55 cm3 g− 1 STP at 298 K and 100 kPa), MIL-100(Fe) (66.12 cm3
g− 1 STP at 298 K and 100 kPa), and DUT-9 (52.00 cm3 g− 1 STP at 298 K
and 100 kPa) [56]. The SF6 adsorption capacities of the Bi-MOFs did not
appear to correlate to the N2 surface areas and porosities of the samples.
This implied that the adsorption of SF6 was not governed by the amount
of available N2 surface area on the Bi-MOF (given that any thermal
expansion of the frameworks has a negligible effect on the diffusion of
SF6). For example, CAU-33 had a lower N2 surface area than CAU-17
however, the SF6 uptake at 100 kPa was found to be slightly higher on
CAU-33. The higher uptake of SF6 on CAU-33 may be due to the

increased presence of pores with dimensions suitable for SF6 sorption
[43]. However, it cannot be ruled out whether the sorption behavior of
SF6 may be influenced by the flexibility of the frameworks, similar to
what was seen for the N2 sorption in UU-200. Such structurally induced
changes were not further studied in this paper. All SF6 isotherms retain a
similar Langmuir-shape for all Bi-MOF structures, apart from CAU-33
(Fig. 5c). The SF6 isotherm of CAU-33 at 293 K shows a clear two-step
isotherm typical for breathing MOFs such as MOFs in the MIL-53 fam­
ily [57–61]. An initial maximum uptake of 13.00 cm3 g− 1 STP is reached

at the phase transition pressure (57 kPa), where the structure changes
from a narrow pore (np) phase to a large pore (lp) phase, with a final
total uptake of 34.84 cm3 g− 1 STP at 100 kPa. The free energy difference
(ΔFhost) between the np and lp phases was calculated to be approxi­
mately 4.5–7 kJ mol− 1 (Table S16), demonstrating that the relative
stability of the two phases were comparable [62,63]. PXRD diffracto­
grams of SF6-saturated CAU-33 also show the presence of new peaks at
lower 2θ (Fig. S6), possibly indicating the presence of a lp-CAU-33 phase
intermixed with np-CAU-33 (although desorption of SF6 may have taken
place while handling the sample). Further studies on the
sorption-induced structural changes of some of these Bi-MOFs could be
interesting for the application of these materials in SF6 sorption and
separation. Low-pressure uptakes of SF6 at 10 kPa on the Bi-MOFs
(Fig. S13) were observed to deviate from the trend shown at 100 kPa,
much in large part due to the breathing phenomenon seen for CAU-33.
CAU-17 (17.71 cm3 g− 1 STP) and UU-200 (16.36 cm3 g− 1 STP) were
found to have the highest uptakes, while SU-101 (11.43 cm3 g− 1 STP)
and CAU-33 (8.52 cm3 g− 1 STP) were shown to have reduced adsorption
capacities of SF6 at the given pressure.
Unlike the sorption behaviors of CO2 and SF6, the CH4 and N2 up­

takes on all Bi-MOFs were found to be relatively low, not exceeding
20.85 cm3 g− 1 STP and 4.93 cm3 g− 1 STP, respectively. The isotherms
for both gases were also found to be linear indicating both a low
adsorption capacity for these gases as well as low affinity. The rest of the
study will therefore focus on the CO2 and SF6 sorption properties of the
Bi-MOFs.
3.4. Gas selectivity and CO2 cycling stability
The calculated gas selectivities for four hypothetical flue gas or
industrially relevant gas mixtures; CH4/N2 (50:50), CO2/CH4 (50:50),
CO2/N2 (15:85), and SF6/N2 (10:90), can be seen in Table 2 and
Fig. 6a–c. The selectivities are calculated from the data obtained at 293
K (closest to room temperaure) only, as the calculated selectivities are
considered to be qualitatively indicative selectivities for real-life appli­
cations. The CO2/N2 IAST selectivities at 100 kPa for all samples were
found to vary between 25 and 35, where SU-101 showed the highest
CO2/N2 selectivity of ~35, which was comparable to that of UU-200
(~35), while CAU-17 had the lowest selectivity of ~25. The CO2/N2
selectivity of UU-200 was also found to be lower than that of porous
sorbents such as Mg(H2gal) (377 at 303 K 100 kPa) [64], Fe(Hgal) (612
at 303 K and 100 kPa) [64], zeolite 13X (981 at 298 K and 100 kPa) [65],
and ZIF-78 (396 at 298 K and 100 kPa) [66]. But comparable to other
MOFs like HKUST-1 (~30 at 273 K and 100 kPa), HNUST-1 (~30 at 298
K and 100 kPa), Co-MOF-74 (~40 at 296 K and 100 kPa), and MIP-202
(~30 at 298 K and 100 kPa) [55]. Although all samples were found to
have Langmuir-shaped CO2 isotherms, indicating an enhanced molecu­
lar interaction between the CO2 molecules and the surface of the pores as
compared to e.g. CH4 and N2, the comparably high N2 uptakes on
CAU-17 and CAU-33 result in an overall lower CO2/N2 selectivity on
these samples. The SF6/N2 IAST selectivities showed a similar trend; the
highest selectivity was observed on UU-200 (~45) followed by SU-101

(~40), and then CAU-17 (~35). These values were comparable to cor­
responding literature values for MOFs such as UiO-66 (~40 at 298 K and
100 kPa) [67] and CAU-17 (~30 at 293 K and 100 kPa) [29]. The SF6/N2
IAST selectivity was not calculated for CAU-33 due to its SF6-induced
flexibility, as the model is typically not applicable for flexible materials
[68]. However, modelling of the SF6 isotherm using a number of
isotherm models (e.g. modified dual-site Langmuir [69]) is possible for
other purposes. It has also been documented in recent literature that the
IAST model has a number of shortcomings; IAST assumes an ideal gas
mixture and is particularly problematic for gases with low interaction
with the sorbent (as integration over a large pressure range is needed if
the selectivity of the sorbent is high). As highlighted by Cheung et al.
[70] and Bjă
ornerbă
ack et al. [71], in order to utilize the IAST model,
binary or multi-component gas adsorption isotherms are needed. We
7


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

Fig. 6. IAST selectivities for hypothetical gas mixtures composed of (a) CO2/CH4 (50:50), (b) CO2/N2 (15:85), (c) SF6/N2 (10:90) and (d) pressure-swing CO2
adsorption cycling on UU-200, CAU-17, CAU-33, and SU-101.

therefore also calculate the CO2 and SF6 selectivity using the Henry’s
law model and the numbers are listed in Table 2 for comparison.
The CH4/N2 and CO2/CH4 selectivities of all Bi-MOFs were found to
be approximately 5. The values were significantly lower compared to

either the CO2/N2 or SF6/N2 selectivities, showing that neither UU-200
nor any of the other investigated Bi-MOFs may have promising gas
separation properties for such gas mixtures.
The isosteric enthalpies of CO2 and SF6 adsorption (-ΔHads) (Fig. 7)
for UU-200 were found to be ~25 kJ mol− 1 (from 6.72 to 38.10 cm3 g− 1
STP CO2 loading) and ~20 kJ mol− 1 (from 8.97 to 23.53 cm3 g− 1 STP
SF6 loading), respectively. The obtained -ΔHads of CO2 was found to be
comparable to that of the other Bi-MOFs; CAU-17 (~25 kJ mol− 1 be­
tween 6.72 and 38.10 cm3 g− 1 STP CO2 loading), CAU-33 (~20 kJ mol− 1
between 6.72 and 38.10 cm3 g− 1 STP CO2 loading) and SU-101 (~20 kJ
mol− 1 between 6.72 and 31.38 cm3 g− 1 STP CO2 loading). The -ΔHads for

all Bi-MOFs, aside from UU-200, can be seen to slightly decrease or
remain constant with an increase in CO2 loading. Which is an expected
trend that can be seen when adsorption sites of lower or comparable
energy become occupied as the degree of loading escalates. The increase
in -ΔHads that is observed for UU-200 was found to likely be related to
the narrow pores in the material. A decrease in internal energy of the
adsorbed phase due to the attractive intermolecular forces between the
CO2 molecule has been proposed to be the cause of this phenomenon
[72]. The calculated -ΔHads of SF6 for UU-200 was however lower than
the corresponding -ΔHads for all other investigated Bi-MOFs; CAU-17
(~35 kJ mol− 1 between 8.97 and 23.53 cm3 g− 1 STP SF6 loading) and
SU-101 (~30 kJ mol− 1 between 6.72 and 16.59 cm3 g− 1 STP SF6
loading). The calculated ranges of all -ΔHads fell below the ranges for
chemisorption, and thus the adsorption of all gases were assumed to be
physisorption-based.

Fig. 7. Isosteric enthalpies of adsorption (-ΔHads) for (a) CO2 and (b) SF6 for UU-200, CAU-17, CAU-33, and SU-101.
8



M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

The physical adsorption of CO2 on the Bi-MOFs was further
demonstrated by the pressure/vacuum-swing CO2 sorption cycling ex­
periments on the Bi-MOFs (Fig. 6d). The CO2 uptake capacities after 5
cycles decreased by less than 0.8% on all Bi-MOFs.
The stability of a MOF is of crucial importance in real-life CO2 cap­
ture and separation applications. As such, the thermal stability of UU200 as well as its physical stability in water was evaluated
(Figs. S7–S8). The framework was found to remain stable up until
approx. 473 K, after which the material was observed to decompose in a
two-step process leading to the formation of Bi2O3 at approx. 673 K.
Additionally, no phase transformation of UU-200 was found to occur
after exposing the material to water for 24 h at room temperature. The
material was also observed to retain its gas adsorption properties when
in the presence of hydrated CO2 gas (~40% RH at 303 K).

Table 3
Comparison of the time required to reach 50% and 90% of the total CO2 uptake
on the Bi-MOFs, respectively.
Sample

50% CO2 uptake (min)

90% CO2 uptake (min)

UU-200

CAU-17
CAU-33
SU-101

2.55
2.77
1.00
1.15

7.53
7.88
3.07
2.37

Thus, the multi-linearity indicates that IP diffusion is not the sole
mechanism limiting the adsorption rate of CO2 in the Bi-MOFs.
Similarly, Boyd plots (Fig. 9b) also show clear deviations from
linearity for all Bi-MOFs, suggesting that film diffusion, i.e. the external
gas film surrounding the particles, and possibly other mechanisms have
an effect on the mass transfer of CO2 in the framework materials. This
correlates well with what was observed from the IP model, i.e. that IP
diffusion does not form the sole rate-limiting step [76]. However,
stronger deviations from linearity were observed for UU-200 and
CAU-17, which may indicate that IP diffusion may play a larger role in
the CO2 adsorption process on CAU-33 and SU-101.
IC diffusional resistance has been shown to play a decisive role in the
adsorption kinetics of many gases in nanoporous framework materials
such as zeolites [77,78]. However, the impact that such diffusional
resistance has on the mass transfer of small gas molecules in MOFs, and
particularly in bismuth-based ones, has not been as well-documented.

Plots from the short- and long-time form of the IC model, Fig. 9c–d,
show good conformity to linearity in the long-time region (Fig. 9d) while
deviations from linearity were observed during the initial stages of
adsorption (Fig. 9c). This suggests that IC diffusion forms the
rate-limiting step towards the end of the adsorption process while
multiple mechanisms limit the adsorption rate during the initial stages.
In order to rule out the influence of heat transfer effects on the diffusion
of CO2 in the materials differential scanning calorimetry (DSC) ther­
mograms were recorded simultaneously with the gravimetric CO2
adsorption profiles on 2.5 mg, 5 mg, and 7.5 mg of sample
(Figs. S21–S24). No significant heat transfer effects related to the
adsorption of CO2 on the MOFs were detected, suggesting that other
effects, such as film-diffusion or external mass-transfer resistance, may
have an influence on the diffusion process.
The diffusion constants, Di r− 2 (Table 4), were calculated from the
Boyd and IC diffusion models using the gravimetric CO2 adsorption
profiles obtained on 2.5 mg, 5 mg, and 7.5 mg of sample. The calculated
CO2 diffusion constants were found to be comparable for all materials,
with CAU-33 and SU-101 showing slightly faster CO2 diffusivities as
compared to the other materials. The diffusion constants were also
found to be within the range for CO2 adsorption in some zeolites (3.64 ×
10− 5 s− 1 to 2.20 × 10− 2 s− 1) [79–82] and MOFs such as mixed-linker
ZIF-7-8s (2.98 × 10− 3 s− 1 to 9.76 × 10− 4 s− 1) [43].

3.5. CO2 adsorption kinetics and estimated diffusivities
The diffusion of CO2 in UU-200 and the selected Bi-MOFs were
studied using a gravimetric method at 303 K (Fig. 8 and Fig. S20). The
adsorption of CO2 was observed to occur rapidly in all materials, with
more than 50% of the total uptake being reached within the first 3 min
(Table 3). Even though the crystallographic pore size of UU-200 was

found be very close to the kinetic diameter of CO2, the observed CO2
adsorption rate on UU-200 still appeared to be both scientific and
possibly industrially relevant. Similar findings have also been observed
for the CO2 adsorption in other microporous materials with comparably
narrow pores [73]. The CO2 adsorption kinetics were found to be slower
in the trimesate-based UU-200 and CAU-17 MOFs than in CAU-33 and
SU-101. This observation could be attributed to both a structural dif­
ference and a difference in the particle size between the samples as can
be seen in the SEM images of the materials (Figs. S10–S11).
Three diffusional models were utilized to further investigate the
adsorption kinetics of CO2, namely, the intraparticle diffusion model
(IP), Boyd’s film diffusion model, and the intracrystalline diffusion
model (IC). The plots obtained from the IP model (Fig. 9a) show clear
multi-linearity corresponding to three regions in the investigated time
interval. Here each region represents a separate step in the adsorption
process, namely: 1) diffusion of the CO2 molecules through the external
gas film surrounding the particles or adsorption through the external
surface of the particles, 2) diffusion of the CO2 molecules through the
pores of the particle (intraparticle diffusion), and 3) slow diffusion of the
molecules through the micropores (at or near equilibrium) [74,75].

4. Conclusions
The CO2 and SF6 adsorption properties of four Bi-MOFs, including a
new bismuth trimesate-based UU-200 MOF, were investigated in detail
in this study. The structure of UU-200, which was determined using a
continuous rotation electron diffraction (cRED) technique, was found to
be different from other existing Bi-trimesate MOFs (e.g. CAU-17). UU200 was found to have good CO2 and SF6 selectivity over N2 with a good
level of CO2 and SF6 uptake at 293 K. The CO2 adsorption kinetics study
suggested that film diffusion and pore diffusion may have governed the
uptake of CO2 at the initial and long-term time scale, respectively. In

summary, we have demonstrated the CO2 and SF6 adsorption properties
of four different Bi-MOFs and investigated the CO2 adsorption kinetics in
detail for the first time. Bi-MOFs showed interesting and selective CO2
and SF6 adsorption properties, which could place them as candidate
greenhouse gas adsorbents for swing adsorption-based applications.

Fig. 8. Gravimetric CO2 adsorption profiles for UU-200, CAU-17, CAU-33, and
SU-100, using approximately 5 mg of sample.
9


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

Fig. 9. (a) Intraparticle diffusion plot, (b) Boyd plot, intracrystalline plot showing the (c) short-time and (d) long-time expression of the model for UU-200, CAU-17,
CAU-33, and SU-101.

Future studies exploring the synthesis conditions of UU-200 (i.e.
different solvents and green synthesis routes), selectivities at different
temperatures, framework flexibility, and the photocatalytic properties
of the MOF may introduce further potential areas of application.

Table 4
Summary of the diffusion constants obtained from Boyd’s film diffusion model
and the intracrystalline diffusion model. The diffusion constant obtained from
the Boyd model are given in brackets due to the data falling outside the appli­
cation range of the model.
Sample


UU200

CAU17

CAU33

SU-101

Approx. sample mass
(mg)

Di r−

2

Declaration of competing interest

(s− 1)

Boyd’s film
diffusion

Intracrystalline diffusion
Short time

Long time

2.5

(7.63 × 10− 4)


5

(4.78 × 10− 4)

7.5

(4.71 × 10− 4)

1.35 ×
10− 4
1.33 ×
10− 4
1.08 ×
10− 4

7.93 ×
10− 4
6.09 ×
10− 4
3.94 ×
10− 4

2.5

(9.66 × 10− 4)

5

(5.88 × 10− 4)


7.5

(3.39 × 10− 4)

1.60 ×
10− 4
1.23 ×
10− 4
6.61 ×
10− 5

1.08 ×
10− 3
6.30 ×
10− 4
3.56 ×
10− 4

2.5

(1.01 × 10− 3)

5

(1.12 × 10− 3)

7.5

(1.26 × 10− 3)


2.31 ×
10− 4
3.49 ×
10− 4
2.75 ×
10− 4

1.16 ×
10− 3
1.21 ×
10− 3
1.30 ×
10− 3

2.5

(1.20 × 10− 3)

5

(9.45 × 10− 4)

7.5

(7.82 × 10− 4)

1.96 ×
10− 4
2.12 ×

10− 4
1.90 ×
10− 4

1.35 ×
10− 3
1.18 ×
10− 3
9.44 ×
10− 4

The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
the work reported in this paper.
Acknowledgments
The authors thank the Swedish Foundation for Strategic Environ­
mental Research (Mistra) (Project Name: Mistra TerraClean, Project
number 2015/31), The Swedish Research Council - Grant no. (OC) 202004029, (MS) 2019-03729, and (TW) 2019-05465 and Swedish Research
Council for Sustainable Development (FORMAS, Grant No. 2018-00651)
ă
ăm and Dr. Francoise M.
for their financial support. Prof. Lars Ohrstr
o
Amombo Noa of Chalmers University of Technology, Sweden are
acknowledged for their input and fruitful discussions.
Appendix A. Supplementary data
CCDC 2103784 contains the supplementary crystallographic data for
this paper. These data can be obtained free of charge from The Cam­
bridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/struc
tures.

Appendix. B Supplementary data
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.micromeso.2021.111548.
10


Microporous and Mesoporous Materials 329 (2022) 111548

M. Åhl´en et al.

References
[27]

[1] S. Mandal, B. Sarkar, N. Bolan, J. Novak, Y.S. Ok, L. Van Zwieten, B.P. Singh, M.
B. Kirkham, G. Choppala, K. Spokas, R. Naidu, Designing advanced biochar
products for maximizing greenhouse gas mitigation potential, Crit. Rev. Environ.
Sci. Technol. 46 (2016) 1367–1401, />10643389.2016.1239975.
[2] R.L. Siegelman, E.J. Kim, J.R. Long, Porous materials for carbon dioxide
separations, Nat. Mater. 20 (2021) 1060–1072, />[3] G. Singh, J. Lee, A. Karakoti, R. Bahadur, J. Yi, D. Zhao, K. AlBahily, A. Vinu,
Emerging trends in porous materials for CO2 capture and conversion, Chem. Soc. Rev.
49 (2020) 4360–4404, />[4] H. Takayabu, CO2 mitigation potentials in manufacturing sectors of 26 countries,
Energy Econ. 86 (2020) 104634, />[5] L.M. Fulton, L.R. Lynd, A. Kă
orner, N. Greene, L.R. Tonachel, The need for biofuels
as part of a low carbon energy future, Biofuel Bioprod. Biorefin. 9 (2015) 476–483,
/>[6] S. Prodinger, R.S. Vemuri, T. Varga, P.B. McGrail, R.K. Motkuri, M.A. Derewinski,
Impact of chabazite SSZ-13 textural properties and chemical composition on CO2
adsorption applications, New J. Chem. 40 (2016) 4375–4385, />10.1039/C5NJ03205A.
[7] O. Cheung, Z. Bacsik, Q. Liu, A. Mace, N. Hedin, Adsorption kinetics for CO2 on highly
selective zeolites NaKA and nano-NaKA, Appl. Energy 112 (2013) 1326–1336,
/>[8] H. Zhang, Z. Wang, X. Luo, J. Lu, S. Peng, Y. Wang, L. Han, Constructing hierarchical

porous carbons with interconnected micro-mesopores for enhanced CO2 adsorption,
Front. Chem. 7 (2019) 919, />[9] S. Jung, Y.-K. Park, E.E. Kwon, Strategic use of biochar for CO2 capture and
sequestration, J. CO2 Util. 32 (2019) 128–139, />jcou.2019.04.012.
[10] M. Ding, R.W. Flaig, H.-L. Jiang, O.M. Yaghi, Carbon capture and conversion using
metal-organic frameworks and MOF-based materials, Chem. Soc. Rev. 48 (2019)
2783–2828, />[11] R. Aniruddha, I. Sreedhar, B.M. Reddy, MOFs in carbon capture-past, present and
future, J. CO2 Util. 42 (2020) 101297, />jcou.2020.101297.
[12] Y. Chen, J. Jiang, A bio-metal-organic framework for highly selective CO2 capture: a
molecular simulation study, ChemSusChem 3 (2010) 982–988, />10.1002/cssc.201000080.
[13] M. Sai Bhargava Reddy, D. Ponnamma, K.K. Sadasivuni, B. Kumar, A.M. Abdullah,
Carbon dioxide adsorption based on porous materials, RSC Adv. 11 (2021)
12658–12681, />[14] H. Li, K. Wang, Y. Sun, C.T. Lollar, J. Li, H.-C. Zhou, Recent advances in gas storage
and separation using metal–organic frameworks, Mater. Today 21 (2018) 108–121,
/>[15] Q. Wang, D. Astruc, State of the art and prospects in metal–organic framework
(MOF)-based and MOF-derived nanocatalysis, Chem. Rev. 120 (2020) 1438–1511,
/>[16] J.-D. Xiao, H.-L. Jiang, Metal-organic frameworks for photocatalysis and
photothermal catalysis, Acc. Chem. Res. 52 (2019) 356–366, />10.1021/acs.accounts.8b00521.
[17] Y. Cui, B. Chen, G. Qian, Lanthanide metal-organic frameworks for luminescent
sensing and light-emitting applications, Coord. Chem. Rev. 273–274 (2014) 76–86,
/>[18] W.P. Lustig, S. Mukherjee, N.D. Rudd, A.V. Desai, J. Li, S.K. Ghosh, Metal-organic
frameworks: functional luminescent and photonic materials for sensing
applications, Chem. Soc. Rev. 46 (2017) 3242–3285, />C6CS00930A.
[19] Y. Sun, L. Zheng, Y. Yang, X. Qian, T. Fu, X. Li, Z. Yang, H. Yan, C. Cui, W. Tan,
Metal-organic framework nanocarriers for drug delivery in biomedical
applications, Nano-Micro Lett. 12 (2020) 103, />[20] T. Qiu, Z. Liang, W. Guo, H. Tabassum, S. Gao, R. Zou, Metal–organic frameworkbased materials for energy conversion and storage, ACS Energy Lett 5 (2020)
520–532, />[21] F. Saraci, V. Quezada-Novoa, P.R. Donnarumma, A.J. Howarth, Rare-earth metalorganic frameworks: from structure to applications, Chem. Soc. Rev. 49 (2020)
7949–7977, />[22] M.A. Alnaqbi, A. Alzamly, S.H. Ahmed, M. Bakiro, J. Kegere, H.L. Nguyen,
Chemistry and applications of s-block metal–organic frameworks, J. Mater. Chem.
9 (2021) 38283854, />ă
[23] Y. Wang, S. Takki, O. Cheung, H. Xu, W. Wan, L. Ohrstră

om, A.K. Inge, Elucidation
of the elusive structure and formula of the active pharmaceutical ingredient
bismuth subgallate by continuous rotation electron diffraction, Chem. Commun. 53
(2017) 7018–7021, />[24] Z. Wang, Z. Zeng, H. Wang, G. Zeng, P. Xu, R. Xiao, D. Huang, S. Chen, Y. He,
C. Zhou, M. Cheng, H. Qin, Bismuth-based metal–organic frameworks and their
derivatives: opportunities and challenges, Coord. Chem. Rev. 439 (2021) 213902,
/>[25] Q.-X. Wang, G. Li, Bi(III) MOFs: syntheses, structures and applications, Inorg.
Chem. Front. 8 (2021) 572–589, />[26] M. Feyand, E. Mugnaioli, F. Vermoortele, B. Bueken, J.M. Dieterich, T. Reimer,
U. Kolb, D. deVos, N. Stock, Automated diffraction tomography for the structure
elucidation of twinned ,sub -micrometercrystals of a highly porous ,catalytically

[28]
[29]

[30]

[31]

[32]

[33]

[34]
[35]
[36]

[37]
[38]

[39]

[40]

[41]
[42]

[43]

[44]
[45]
[46]
[47]
[48]
[49]

[50]

11

active bismuth metal -organicframework, Angew. Chem. Int. Ed. 51 (2012)
10373–10376, />M. Savage, S. Yang, M. Suyetin, E. Bichoutskaia, W. Lewis, A.J. Blake, S.A. Barnett,
M. Schră
oder, A novel bismuth-based metal-organic framework for high volumetric
methane and carbon dioxide adsorption, Chem. Eur J. 20 (2014) 80248029,
/>M. Kă
oppen, O. Beyer, S. Wuttke, U. Lüning, N. Stock, Synthesis, functionalisation
and post-synthetic modification of bismuth metal-organic frameworks, Dalton
Trans. 46 (2017) 8658–8663, />M. Kă
oppen, A. Dhakshinamoorthy, A.K. Inge, O. Cheung, J. ngstră
om, P. Mayer,
N. Stock, Synthesis, transformation, catalysis, and gas sorption investigations on

the bismuth metal-organic framework CAU-17, Eur. J. Inorg. Chem., 2018 (2018)
34963503, />M. Kă
oppen, V. Meyer, J. ngstră
om, A.K. Inge, N. Stock, Solvent-dependent
formation of three new Bi-metal–organic frameworks using a tetracarboxylic acid,
Cryst. Growth Des. 18 (2018) 4060–4067, />cgd.8b00439.
L. Robison, L. Zhang, R.J. Drout, P. Li, C.R. Haney, A. Brikha, H. Noh, B.L. Mehdi,
N.D. Browning, V.P. Dravid, Q. Cui, T. Islamoglu, O.K. Farha, A bismuth
metal–organic framework as a contrast agent for X-ray computed tomography, ACS
Appl. Bio Mater. 2 (2019) 1197–1203, />E. Svensson Grape, J.G. Flores, T. Hidalgo, E. Martớnez-Ahumada, A. Gutierreză
Alejandre, A. Hautier, D.R. Williams, M. OKeeffe, L. Ohrstră
om, T. Willhammar,
P. Horcajada, I.A. Ibarra, A.K. Inge, A robust and biocompatible bismuth ellagate
MOF synthesized under green ambient conditions, J. Am. Chem. Soc. 142 (2020)
1679516804, />A.K. Inge, M. Kă
oppen, J. Su, M. Feyand, H. Xu, X. Zou, M. O’Keeffe, N. Stock,
Unprecedented topological complexity in a metal-organic framework constructed
from simple building units, J. Am. Chem. Soc. 138 (2016) 1970–1976, https://doi.
org/10.1021/jacs.5b12484.
F. Ye, Z.-X. Wei, J.-F. Song, X.-H. Wu, P. Yue, Synthesis, characterization, and
photocatalytic properties of bismuth (III)-benzene-1,3,5-tricarboxylate, Z. Anorg.
Allg. Chem. 643 (2017) 669–674, />G. Wang, Y. Liu, B. Huang, X. Qin, X. Zhang, Y. Dai, A novel metal-organic
framework based on bismuth and trimesic acid: synthesis, structure and properties,
Dalton Trans. 44 (2015) 16238–16241, />Y. Sun, N. Zhang, Q.L. Guan, C.H. Liu, B. Li, K.Y. Zhang, G.H. Li, Y.H. Xing, F.
Y. Bai, L.X. Sun, Sensing of Fe3+ and Cr2O72– in water and white light: synthesis,
characterization, and fluorescence properties of a crystalline bismuth-1,3,5benzenetricarboxylic acid framework, Cryst. Growth Des. 19 (2019) 7217–7229,
/>M. Sanchez-Sala, O. Vallcorba, C. Domingo, J.A. Ayll´
on, Acetic acid as a solvent for
the synthesis of metal–organic frameworks based on trimesic acid, Polyhedron 170
(2019) 458462, />M.O. Cichocka, J. ngstră

om, B. Wang, X. Zou, S. Smeets, High-throughput
continuous rotation electron diffraction data acquisition via software automation,
J. Appl. Crystallogr. 51 (2018) 1652–1661, />S1600576718015145.
W. Kabsch, Xds, Acta crystallogr, D Biol. Crystallogr. 66 (2010) 125–132, https://
doi.org/10.1107/S0907444909047337.
M.C. Burla, R. Caliandro, B. Carrozzini, G.L. Cascarano, C. Cuocci, C. Giacovazzo,
M. Mallamo, A. Mazzone, G. Polidori, Crystal structure determination and
refinement via SIR2014, J. Appl. Crystallogr. 48 (2015) 306–309, />10.1107/s1600576715001132.
G.M. Sheldrick, Crystal structure refinement with SHELXL, Acta Crystallogr. C
Struct. Chem. 71 (2015) 3–8, />E. Svensson Grape, H. Xu, O. Cheung, M. Calmels, J. Zhao, C. Dejoie, D.
M. Proserpio, X. Zou, A.K. Inge, Breathing metal–organic framework based on
flexible inorganic building units, Cryst. Growth Des. 20 (2019) 320–329, https://
doi.org/10.1021/acs.cgd.9b01266.
M. Åhl´en, A. Jaworski, M. Strømme, O. Cheung, Selective adsorption of CO2 and SF6
on mixed-linker ZIF-7–8s: The effect of linker substitution on uptake capacity and
kinetics, Chem. Eng. J. 422 (2021) 130117, />cej.2021.130117.
ˇ
ˇ
B. Zdravkov, J. Cerm´
ak, M. Sefara,
J. Janků, Pore classification in the
characterization of porous materials: a perspective, Cent. Eur. J. Chem. 5 (2007)
385–395, />D.W. Breck, Zeolite Molecular Sieves: Structure, Chemistry and Use, John Wiley &
Sons, New York, 1974.
A.F. Ismail, K.C. Khulbe, T. Matsuura, in: Gas Separation Membranes - Polymeric
and Inorganic, first ed., Springer International Publishing, New York, 2015.
S.S. Han, W.A. Goddard, Metal-organic frameworks provide large negative thermal
expansion behavior, J. Phys. Chem. C 111 (2007) 15185–15191, />10.1021/jp075389s.
D. Dubbeldam, K.S. Walton, D.E. Ellis, R.Q. Snurr, Exceptional negative thermal
expansion in isoreticular metal-organic frameworks, Angew. Chem. Int. Ed. 46

(2007) 4496–4499, />C. Schneider, D. Bodesheim, M.G. Ehrenreich, V. Crocell`
a, J. Mink, R.A. Fischer, K.
T. Butler, G. Kieslich, Tuning the negative thermal expansion behavior of the metalorganic framework Cu3BTC2 by retrofitting, J. Am. Chem. Soc. 141 (2019)
10504–10509, />M.J. Cliffe, J.A. Hill, C.A. Murray, F.-X. Coudert, A.L. Goodwin, Defect-dependent
colossal negative thermal expansion in UiO-66(Hf) metal-organic framework, Phys.
Chem. Chem. Phys. 17 (2015) 11586–11592, />C5CP01307K.


M. Åhl´en et al.

Microporous and Mesoporous Materials 329 (2022) 111548

[51] J.D. Evans, J.P. Dürholt, S. Kaskel, R. Schmid, Assessing negative thermal
expansion in mesoporous metal–organic frameworks by molecular simulation,
J. Mater. Chem. 7 (2019) 24019–24026, />[52] B. Garai, V. Bon, A. Efimova, M. Gerlach, I. Senkovska, S. Kaskel, Reversible
switching between positive and negative thermal expansion in a metal–organic
framework DUT-49, J. Mater. Chem. 8 (2020) 20420–20428, />10.1039/D0TA06830F.
[53] H.-L. Zhou, Y.-B. Zhang, J.-P. Zhang, X.-M. Chen, Supramolecular-jack-like guest in
ultramicroporous crystal for exceptional thermal expansion behaviour, Nat.
Commun. 6 (2015) 6917, />[54] H.-L. Zhou, R.-B. Lin, C.-T. He, Y.-B. Zhang, N. Feng, Q. Wang, F. Deng, J.-P. Zhang,
X.-M. Chen, Direct visualization of a guest-triggered crystal deformation based on a
flexible ultramicroporous framework, Nat. Commun. 4 (2013) 2534, https://doi.
org/10.1038/ncomms3534.
[55] Z. Li, P. Liu, C. Ou, X. Dong, Porous metal–organic frameworks for carbon dioxide
adsorption and separation at low pressure, ACS Sustain. Chem. Eng. 8 (2020)
15378–15404, />[56] D.K.J.A. Wanigarathna, J. Gao, B. Liu, Metal organic frameworks for adsorptionbased separation of fluorocompounds: a review, Mater. Adv. 1 (2020) 310–320,
/>[57] F. Millange, N. Guillou, R.I. Walton, J.-M. Gren`
eche, I. Margiolakid, G. F´erey,
Effect of the nature of the metal on the breathing steps in MOFs with dynamic
frameworks, Chem. Commun. (2008) 4732–4734, />B809419E.

[58] T. Loiseau, C. Serre, C. Huguenard, G. Fink, F. Taulelle, M. Henry, T. Bataille,
G. F´erey, A rationale for the large breathing of the porous aluminum terephthalate
(MIL-53) upon hydration, Chemistry 10 (2004) 1373–1382, />10.1002/chem.200305413.
[59] C. Volkringer, T. Loiseau, N. Guillou, G. F´
erey, E. Elkaïmc, A. Vimont, XRD and IR
structural investigations of a particular breathing effect in the MOF-type gallium
terephthalate MIL-53(Ga), Dalton Trans. (2009) 2241–2249, />10.1039/B817563B.
[60] C. Serre, F. Millange, C. Thouvenot, M. Nogu`es, G. Marsolier, D. Louă
er, G. F
erey,
Very large breathing effect in the first nanoporous chromium(III)-based solids: MIL-53 or
III
Cr (OH)⸱{O2C-C6H4-CO2}⸱{HO2C-C6H4-CO2H}x⸱H2Oy, J. Am. Chem. Soc. 124
(2002) 13519–13526, />[61] J.P.S. Mowat, V.R. Seymour, J.M. Griffin, S.P. Thompson, A.M.Z. Slawin, D. FairenJimenez, T. Düren, S.E. Ashbrooka, P.A. Wright, A novel structural form of MIL-53
observed for the scandium analogue and its response to temperature variation and CO2
adsorption, Dalton Trans. 41 (2012) 3937–3941, />C1DT11729G.
[62] Z. Wang, S.M. Cohen, Modulating metal-organic frameworks to breathe: a
postsynthetic covalent modification approach, J. Am. Chem. Soc. 131 (2009)
16675–16677, />[63] F.-X. Coudert, M. Jeffroy, A.H. Fuchs, A. Boutin, C. Mellot-Draznieks,
Thermodynamics of guest-induced structural transitions in hybrid organicinorganic frameworks, J. Am. Chem. Soc. 130 (2008) 14294–14302, https://doi.
org/10.1021/ja805129c.
[64] N. Heymans, S. Bourrelly, P. Normand, E. Bloch, H. Mkhadder, L. Cooper,
M. Gorman, I. Bouzidi, Nathalie Guillou, G. De Weireld, T. Devic, Small-pore
gallates MOFs for environmental applications: Sorption behaviors and structural
elucidation of their high affinity for CO2, J. Phys. Chem. C 124 (2020) 3188–3195,
/>[65] A. Kumar Rajagopalan, A.M. Avila, A. Rajendran, Do adsorbent screening metrics
predict process performance? A process optimisation based study for post-combustion

[66]
[67]


[68]
[69]
[70]
[71]
[72]

[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]

12

capture of CO2, Int. J. Greenh. Gas Control. 46 (2016) 76–85, />10.1016/j.ijggc.2015.12.033.
Y.-S. Bae, R.Q. Snurr, Development and evaluation of porous materials for carbon
dioxide separation and capture, Angew. Chem. Int. Ed. 50 (2011) 11586–11596,
/>M.-B. Kim, S.-J. Lee, C.Y. Lee, Y.-S. Bae, High SF6 selectivities and capacities in
isostructural metal-organic frameworks with proper pore sizes and highly dense
unsaturated metal sites, Microporous Mesoporous Mater. 190 (2014) 356–361,
/>G. Fraux, A. Boutin, A.H. Fuchs, F.-X. Coudert, On the use of the IAST method for
gas separation studies in porous materials with gate-opening behavior, Adsorption
24 (2018) 233–241, />P. Mishra, H.P. Uppara, B. Mandal, S. Gumma, Adsorption and separation of carbon
dioxide using MIL-53(Al) metal-organic framework, Ind. Eng. Chem. Res. 53

(2014) 19747–19753, />O. Cheung, Z. Bacsik, N. Fil, P. Krokidas, D. Wardecki, N. Hedin, Selective
Adsorption of CO2 on zeolites NaK-ZK-4 with Si/Al of 1.8-2.8, ACS Omega 5 (2020)
2537125380, />F. Bjă
ornerback, D. Bernin, N. Hedin, Microporous humins synthesized in
concentrated sulfuric acid using 5-hydroxymethyl furfural, ACS Omega 3 (2018)
8537–8545, />S. Hlushak, Heat of adsorption, adsorption stress, and optimal storage of methane
in slit and cylindrical carbon pores predicted by classical density functional theory,
Phys. Chem. Chem. Phys. 20 (2018) 872–888, />C7CP06591D.
O. Cheung, N. Hedin, Zeolites and related sorbents with narrow pores for CO2
separation from flue gas, RSC Adv. 4 (2014) 14480–14494, />10.1039/c3ra48052f.
F.-C. Wu, R.-L. Tseng, R.-S. Juang, Initial behavior of intraparticle diffusion model
used in the description of adsorption kinetics, Chem. Eng. J. 153 (2009) 1–8,
/>H.K. Boparai, M. Joseph, D.M. O’Carroll, Kinetics and thermodynamics of
cadmium ion removal by adsorption onto nano zerovalent iron particles, J. Hazard
Mater. 186 (2011) 458–465, />C. Yao, T. Chen, A film-diffusion-based adsorption kinetic equation and its
application, Chem. Eng. Res. Des. 119 (2017) 87–92, />cherd.2017.01.004.
D.M. Ruthven, Principles of Adsorption and Adsorption Processes, John Wiley Sons
Inc., New Jersey, 1984.
J. Kă
arger, D.M. Ruthven, D.N. Theodorou, Diffusion in Nanoporous Materials,
Wiley-VCH Verlag GmbH, 2012.
H.-W. Ahn, J.-H. Moon, S.-H. Hyun, C.-H. Lee, Diffusion mechanism of carbon
dioxide in zeolite 4A and CaX pellets, Adsorption 10 (2004) 111–128, https://doi.
org/10.1023/B:ADSO.0000039867.14756.ac.
N. Haq, D.M. Ruthven, Chromatographic study of sorption and diffusion in 4A
zeolite, J. Colloid Interface Sci. 122 (1986) 154–163, />0021-9797(86)90077-9.
Y. Hayrettin, D.M. Ruthven, Diffusion of CO2 in 4A and 5A zeolite crystals, J. Colloid
Interface Sci. 74 (1980) 186–195, />V.K. Kaushik, R.P. Vijayalaksmi, N.V. Choudary, S.G.T. Bhat, XPS studies on cation
exchanged zeolite A, Microporous Mesoporous Mater. 51 (2002) 139–144, https://
doi.org/10.1016/S1387-1811(01)00473-5.




×