Vous êtes sur la page 1sur 11

Sorption equilibrium and kinetics of CO

2
on clay minerals from
subcritical to supercritical conditions: CO
2
sequestration at nanoscale
interfaces
Pil Rip Jeon
a
, Jiwon Choi
a
, Tae Sup Yun
b
, Chang-Ha Lee
a,
a
Department of Chemical and Biomolecular Engineering, Yonsei University, Seoul, South Korea
b
Department of Civil and Environmental Engineering, Yonsei University, Seoul, South Korea
h i g h l i g h t s
v The sorption isotherm and rate of
high pressure CO
2
among clay
minerals were studied.
v Structural change of clay minerals
was observed after supercritical CO
2
sorption.
v Excess sorbed amount of CO
2
on clay
mineral showed a maximum near
critical pressure.
v The absolute sorbed amount changed
negligibly after the critical density of
CO
2
.
g r a p h i c a l a b s t r a c t
a r t i c l e i n f o
Article history:
Received 3 April 2014
Received in revised form 16 June 2014
Accepted 23 June 2014
Available online 1 July 2014
Keywords:
Clay mineral
Carbon dioxide
Sorption capacity
Sorption rate
Supercritical condition
a b s t r a c t
CO
2
sequestration in geological formations has attracted attention as a promising method to reduce
anthropogenic CO
2
emission. CO
2
sorption at nanoscale interfaces of clay minerals were studied from
subcritical to supercritical conditions because clay minerals are a constituent of various rocks such as
a cap rock, reservoir rock and coal mineral matter. The sorption capacity and kinetics of CO
2
on montmo-
rillonite, illite, and sepiolite were measured by a gravimetric method. Sepiolite had the highest sorption
capacity at all experimental conditions. After high CO
2
pressure sorption, the desorption isotherm on
montmorillonite showed signicant hysteresis, but the hysteresis on illite and sepiolite was relatively
weak. The excess sorption isotherms of all clay minerals showed a maximum near the critical pressure
and the absolute sorption isotherms approached the saturation over the critical density value of CO
2
.
The surface area changes of clay minerals by supercritical CO
2
sorption were observed by comparing
the N
2
sorption isotherms between the raw material and post-experiment sample. The CO
2
sorption rates
on clay minerals were within a single order of magnitude (10
8
m
2
/s). The results at nanoscale interfaces
can contribute to understanding the sorption capacity and sealing integrity of sedimentary rocks in CO
2
geological storage.
2014 Elsevier B.V. All rights reserved.
1. Introduction
As CO
2
is a major contributor to the greenhouse effect,
capturing and sequestering CO
2
has emerged as a major global
issue. Storage in geological formations is under investigation as
method to reduce anthropogenic CO
2
emission. A signicant
advantage of sequestration is its potential for alleviating environ-
mental damage to the atmosphere without noticeable change to
current lifestyles [1]. Potential sites for sequestration are saline
aquifers, depleted gas and oil reservoirs or unminable coal seams
[24]. The capacity estimated for these reservoirs illustrates that
http://dx.doi.org/10.1016/j.cej.2014.06.090
1385-8947/ 2014 Elsevier B.V. All rights reserved.

Corresponding author. Tel.: +82 2 2123 2762; fax: +82 2 312 6401.
E-mail address: leech@yonsei.ac.kr (C.-H. Lee).
Chemical Engineering Journal 255 (2014) 705715
Contents lists available at ScienceDirect
Chemical Engineering Journal
j our nal homepage: www. el sevi er . com/ l ocat e/ cej
the geological sequestration of CO
2
may have sufcient potential.
As a result, studies have been conducted to conrm the efciency
and practicability of geological sequestration.
Although some benets of CO
2
sequestration in geological stor-
age have attracted attention, risks also occur due to the possibility
of leakage of the CO
2
to the subsurface or atmosphere. Leakage
through cap rocks may occur in several ways: rapid leakage by
damage of the well casing, long-term leakage controlled by the
capillary sealing efciency [57], and diffusive migration of the
dissolved gas through the water-saturated pore space. Therefore,
another common issue of CO
2
capture and sequestration is the
sealing efciency of cap rocks above the potential CO
2
storage res-
ervoir to assess the leakage risk and leakage rates for site approval,
public acceptance, and awarding of credits for stored CO
2
quanti-
ties [8]. To overcome these uncertainties, many studies have been
conducted to understand the CO
2
waterrock interaction in aqui-
fers for the quantitative assessment of CO
2
capacity and integrity
[1,811].
Several studies have been conducted over the past decade to
estimate the CO
2
storage capacity and diffusion rate in sedimen-
tary basins [12,13]. A sedimentary rock consists of grain, cement,
and porosity. When injected into sedimentary basins, high pres-
sure CO
2
is captured in an intergranular porosity of sedimentary
rocks. Simultaneously, high pressure CO
2
penetrates the intragran-
ular porosity of grains (sand grains, fossils, and ooids), cement
(chemically-precipitated mineral matter), and matrix (clay-sized
sediment). The grains, cement, and matrix have relatively large
pores (macropores and mesopores) between crystals and small
pores (micropores) in the crystals (Fig. 1). Therefore, after CO
2
mol-
ecules move along the diffusion path of macropores and mesop-
ores, they reach the micropores with a relatively large surface
area and surface afnity. The micropores can offer a large sorption
capacity and sorptive stabilization to CO
2
.
The mechanisms of CO
2
sorption on sedimentary strata are
complex and the sedimentary environment is diverse. The proper-
ties of the sedimentary rock depend on pore features such as depo-
sitional setting and original mineralogy, sediment compaction,
burial depth and diagenetic alteration, and deformation history
[14]. And the sorption behaviors are also inuenced by the proper-
ties of the pore uid, such as density and interfacial tension (IFT).
Therefore, the mechanism of dry CO
2
sorption on the sediment
needs to be understood rst.
Clay minerals are major constituent of various rocks. Some clay
minerals, like mica and the smectite group, can constitute up to
30% of reservoir rocks and cap rocks [14]. Therefore, understanding
the sorption mechanism of CO
2
at nanoscale interfaces of clay min-
erals can partly contribute to evaluating sorption capacity and
sealing efciency of sedimentary rocks in CO
2
geological storage.
Furthermore, once injected, high pressure CO
2
can spread through-
out the geological strata or leak to other strata. This will lead to a
decrease in the injected CO
2
pressure and the CO
2
sorption capacity
may vary from that at the injected pressure. It was reported that
the desorption behavior of CO
2
from coal may be different from
the sorption behavior due to the hysteresis [15,16]. In addition,
the physical structure of coal with a high amount of ash can be
affected by high pressure CO
2
[15,16]. However, it is not clear
whether the change in physical properties stems from coal struc-
ture itself or mineral matter in the coal. Therefore, it is needed to
study the sorption and desorption of CO
2
on clay minerals.
In this study, the sorption and desorption characteristics of sub-
critical to supercritical CO
2
on dry clay minerals were measured by
a gravimetric method. Montmorillonite, illite, and sepiolite were
selected as representative clay minerals. The sorption isotherms
were measured up to 120 bar at 318 K and 328 K and the desorp-
tion isotherms were measured by decreasing CO
2
pressure from
the applied maximum sorption pressure. The sorption capacity
and rate of CO
2
were compared among the clay minerals. Struc-
tures of the clay minerals before and after the high-pressure sorp-
tion/desorption experiments were compared to detect the changes
resulting from exposure to CO
2
.
2. Experimental
2.1. Sample description
The clay minerals used in this study were obtained from the
Clay Minerals Society. Considering the structural difference, three
kinds of clay mineral, montmorillonite, illite, and sepiolite (CMS
source clay SCa-3, IMt-1, and SepSp-1, respectively), were chosen
as the sorbent to assess the sorption capacity of CO
2
.
Montmorillonite is a sheet structure composed of a layer of
octahedral aluminum oxide between two layers of tetrahedral sil-
icon oxides. The spaces between the layered sheets are available
for sorption and substitution of exchangeable cations. It may be
possibly swelling due to high pressure sorption. Illite is the chief
constituent in many shales. Although the layered structure of illite
is very similar to that of montmorillonite, the basal spacing is xed
and non-expansive due to existence of potassium ions between
layers [14]. Sepiolite contains two-dimensional tetrahedral sheets,
however they differ from other layer silicates in that they lack con-
tinuous octahedral sheets and contain a ribbon structure. Although
sepiolite appears to be a metastable phase with respect to saponite
and montmorillonite [17], it is a non-swelling mineral. The formu-
lae and structures of the clay mineral are summarized in Table 1.
Using N
2
sorption/desorption analysis at 77 K (Quantachrome
Instrument, Autosorb-iQ MP), the structural properties of the three
mineral samples were measured. The samples were gently ground
to a particle size of 150500 lm. Before the experiment, the
Fig. 1. Schematics of components and CO
2
sorption of sedimentary rocks.
Table 1
Structure and formula of montmorillonite, illite, and sepiolite.
Formula
a
Structure
Montmorillonite (Na,Ca)
0.33
(Al,Mg)
2
(Si
4
O
10
)(OH)
2
nH2O
Sheet structure of octahedral
aluminum oxide and
tetrahedral silicon oxide
Illite (K,H
3
O)(Al,Mg,Fe)
2
(Si,Al)
4
O
10
[(OH)
2
,(H
2
O)]
Phyllosilicate or layered
alumino-silicate constituted by
the repetition of tetrahedron
octahedron tetrahedron
(TOT) layers
Sepiolite Mg
4
Si
6
O
15
(OH)
2
6H
2
O A hydrous magnesium silicate
whose individual particles have
a needle-like morphology
a
Reference: http://www.clays.org/.
706 P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715
sample was dried under vacuum at 378 K for more than 12 h. Since
the sample might have taken up atmospheric moisture when
placed on the sample basket, vacuum drying was conducted again
in a sorption system at 373 K for 3 h. In the study, the DR (Dubi-
ninRadushkevich) method and BJH (BarrettJoynerHalenda)
method were applied to estimate the micropore surface area and
mesopore surface area, respectively. The DR method was used at
the P/P
0
range up to 0.1, and the BJH method was applied for the
P/P
0
range of greater than 0.35. In the case of BET (Brunauer
EmmettTeller) surface area, it was estimated in the P/P
0
range
of 0.050.3.
The specic surface areas of montmorillonite, illite, and sepio-
lite were 87.77, 23.25, and 309.56 m
2
/g, respectively. The surface
area of montmorillonite was reported as ~800 m
2
/g when it was
measured by the H
2
O molecule [18]. However, it was also pointed
out that it was much greater than the N
2
BET surface area because
of the polarity of H
2
O. In the study, N
2
BET surface area was mea-
sured for the comparison of the before/after experimental samples
because CO
2
is non-polar molecules. In addition, the BET surface
area measured in the study was similar to that in previous reports
[19,20].
The nitrogen sorption isotherm at 77 K conrmed microporos-
ity and mesoporosity of the clay mineral. Since different models
were used for micropore and mesopore and the applied pressure
range (P/P
0
) of each model was different, the summation of surface
area of micro and mesopore was different from the BET surface
area in Table 2. In the study, the structural property was used to
conrm the relative change of the samples after exposure to high
pressure CO
2
. As shown in Fig. 2, montmorillonite exhibited a
strong hysteresis in a P/P
0
range of 0.51.0. Considering the scale
of y axis, certain level of hysteresis was observed for the sepiolite
in the P/P
0
range from 0.8 to 1.0. Very weak hysteresis of illite
was observed in the P/P
0
range from 0.4 to 1.0.
2.2. Sorption measurements
The sorption experiments on the clay minerals were performed
by the gravimetric method using a magnetic suspension balance
(MSB; Rubotherm, Bochum, Germany) shown in Fig. 3. The MSB
consists of a basket containing the sorbent, a Ti-sinker, and a per-
manent magnet suspended by an electromagnet inside a sorption
cell that separates the sensitive microbalance from the sample
and measuring atmosphere.
The set-up of the Ti-sinker allowed the density of the bulk uid
to be measured in situ. The pressure was measured with an electri-
cal pressure gauge (GE Sensing, PMP 4010) with an accuracy of
0.08% at the full scale value. The balance was maintained at a con-
stant temperature using a heating jacket and the temperature was
measured using a RTD (Thermosensor GmbH, Pt100 RTD) with an
accuracy of 0.05 K. The weight of the sample was measured with
an accuracy of 0.01 mg. Less than 1 g of the sample was applied to
each experiment. With this method, the relative errors do not
accumulate with increasing sorption pressure. However, a stable
reading was acquired over a long time because the balance was
very sensitive to any temperature gradients inside the sorption cell
[21].
Both the amount of excess sorbed and the density of the bulk
uid were measured by a general gravimetric method [22,23]
and the method was described in the Appendix.
The experimental results are generally reported in terms of the
molar excess sorption (n
ex
), which is dened as moles of sorbate
per unit mass of sorbent.
n
ex
=
m
ex
(q
b
; T)
Mm
0
sorbent
(1)
where M is the molar mass of sorbing gas.
The sorption experiments of CO
2
on each clay mineral were per-
formed at 318 and 328 K. After clay mineral activation and He
purge, the MSB loaded with clay mineral was outgassed at
0.05 bar. The pressure was then increased in a step-wise manner
up to 120 bar. Therefore, CO
2
sorption was performed from the
gaseous phase to the supercritical uid. The corresponding uptake
of sorption with time was monitored at a xed pressure condition
until the sorption reached equilibrium. To measure the equilibrium
Table 2
Physical properties of clay minerals.
BET surface area
(m
2
/g)
Micropore surface area
(m
2
/g)
Mesopore surface area
(m
2
/g)
e q
p
(g/cm
3
)
Montmorillonite 87.77 80.97 73.49 0.0270 2.787
Illite 23.25 18.46 30.56 0.0368 3.338
Sepiolite 309.6 300.7 172.2 0.2777 2.627
Fig. 2. Nitrogen sorption isotherms of montmorillonite, illite, and sepiolite at 77 K.
Fig. 3. Schematic diagram of gravimetric sorption equipment.
P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715 707
value for sorption and desorption, 1600 min was applied to each
experimental run at rst. Then, it was decided that an equilib-
rium was reached if the value was not changed in the range of
10
4
g/h.
After sorption equilibrium at 120 bar, the experiment to mea-
sure the desorption isotherm was conducted by decreasing
pressure from the MSB. After the sorption/desorption experiments,
N
2
sorption/desorption analysis at 77 K was performed to compare
structural properties between the clay minerals before and after
exposure to high pressure CO
2
. Powder X-ray diffraction patterns
for randomly oriented montmorillonite were recorded using a
Rigaku D/MAX 2500 V X-ray diffractometer with a copper target
operating at 40 kV and 100 mA. The diffractograms were taken
with a scanning rate of 0.02 degrees/s.
2.3. Kinetic uptake and diffusion model
Information regarding the CO
2
stabilization time at nanoscale
interfaces is important for understanding the geological CO
2
storage.
Transport processes, such as molecular diffusion, Knudsen dif-
fusion, surface diffusion, and diffusion in a micropore, make signif-
icant contributions to CO
2
sorption on clay minerals. In addition,
swelling or structural change cause the modeling of the kinetic
process to be very difcult and complex. In this study, the simplest
approach, Ficks diffusion law for homogenous materials as a single
porous particle [2426], was applied for clay mineral.
The following mass balance equation can be obtained by per-
forming a mass balance around a thin shell element in the spheri-
cal particle [27].
e
@C
@t
(1 e)
@q
@t
= eD
p
1
r
2
@
@r
r
2
@C
@r

(1 e)D
s
1
r
2
@
@r
r
2
@q
@r

(2)
where e is the transport void fraction of the particle, C is the gas
concentration, q is the concentration in the sorbed phase, D
p
is
the pore diffusivity, D
s
is the surface diffusivity, and r is the particle
radius. In the study, the voidage is the transport void fraction, which
is the void space for the transport. It implies that the transport of
free molecules is available for macropore and mesopore [27]. Since
the porosity can be known as the macropore porosity, the transport
void fraction, e, was measured by a Hg porosimeter (Quantachrome,
Poremaster). And the pore diameter in the range of 3.5 nm
1000 lm was used to estimate the void fraction.
In the study, it took a long time to reach equilibrium due to very
slow diffusion. Therefore, the heat transfer was assumed to be suf-
ciently rapid compared to the sorption rate, so that temperature
gradients both through the particle and between the particle and
surrounding uid were negligible [28]. Assuming that the sorption
in clay mineral is controlled by apparent diffusion, Eq. (2) could be
written in terms of only the gas phase concentration.
@C
@t
= D
app
\
2
C (3)
The parameter D
app
is the apparent diffusivity and the solution
of the equation is
M
t
M
eq
= 1
6
p
2
X

n=1
exp
D
app
n
2
p
2
t
r
2

(4)
where M
t
is the gas uptake at time t, M
eq
is the gas uptake at equi-
librium, and D
app
can be dened as follows:
D
app
=
eD
p
e (1 e)f
/
(C)
(5)
f
/
(C) =
dq
dC
= q
p
RT
@q
@P

T
(6)
D
app
is affected by the sorption amount at specic P and T in
Eq.(6). In addition, D
p
is related to the diffusion mechanism in that
molecular diffusion and Knudsen diffusion should be considered.
1
D
p
= s
1
D
k

1
D
m

(7)
D
k
= 4850d
T
M
i
1
2
(8)
D
m
=
2T
3=2
3p
3=2
r
2
A
P
j
3
N
M
i

1=2
(9)
where q
P
is the particle density, R is the gas constant, s is the tor-
tuosity, d is the pore diameter, T is the absolute temperature, N is
Avogadros number, P is the system pressure, j is the Boltzmann
constant, and r
A
is the Lennard-Jones diameter of the spherical mol-
ecules A, and d is the mesopore diameter. The void fraction (ratio of
void volume of inter-granules to total volume) and density of the
clay mineral are summarized in Table 2.
Eq. (4) contains a rapidly converging series with the apparent
diffusion coefcient D
app
and time t for a given particle size [26].
To get the apparent diffusion coefcient, the particle size was cal-
culated from the total volume of the sample. A plot of
ln(1 M
t
=M

) against time t is linear in the region of M


t
=M

>
0:5, thereby allowing the diffusion coefcient to be calculated from
the gradient with 95% accuracy with respect to the theoretical
value.
3. Results and discussion
3.1. Sorption on clay minerals
Fig. 4(a) and (b) show the excess sorbed amount of CO
2
on
montmorillonite measured at 318 and 328 K. The excess isotherms
of CO
2
at both temperatures increase gradually with pressure and
then decrease after passing the critical pressure. The sorption iso-
therm at 318 K is slightly higher than that at 328 K over the entire
pressure range. The 318 K excess sorption isotherm has its maxi-
mum close to the critical pressure while the maximum of the
328 K isotherm lies near 90 bar.
Desorption isotherms at both temperatures showed positive
hysteresis over the sorption isotherm and the hysteresis was
slightly greater at 328 K. The maximum excess sorbed amount at
328 K was greater than that at 318 K in the desorption isotherm.
The maximum values of the desorption isotherm at both tempera-
tures were observed at the gas phase (subcritical condition).
It was reported that the swelling of the sorbents should be
accounted for the accurate sorption amount because the swollen
material occupies a greater free volume [16,29,30]. Even though
the buoyancy effect by the CO
2
pressure on the sorption amount
was calibrated in the gravimetric system, the necessary correction
for the occupied volume of sample could not be made [29]. On the
other hand, it was pointed out in the studies of coal-CO
2
systems
that the level of swelling increases with the CO
2
pressure and the
coal swelling has to be considered for the accurate measurement
of CO
2
sorption [16,29].
Montmorillonite has a sheet structure of octahedral aluminum
oxide and tetrahedral silicon oxide as shown in Table 1. As men-
tioned in the previous studies, in Fig. 4(a) and (b), swelling may
contribute to the hysteresis [31] and the higher hysteresis seems
to result from more swelling at a higher temperature [32]. How-
ever, the degree of the hysteresis in montmorillonite was much
708 P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715
smaller than that in coal [15]. The excess sorbed CO
2
on illite is also
presented in Fig. 4(c) and (d). The sorption isotherm showed
greater capacity at 318 K than at 328 K at subcritical conditions.
The isotherm of illite reaches a maximum in the range of 40
50 bar and decreases smoothly. It then decreases abruptly in the
supercritical condition and even reached negative values. A nega-
tive excess sorbed CO
2
can be expected if the density of the sorbed
phase is less than that of the free gas phase [33].
Compared to montmorillonite, the excess sorbed amounts on
illite, including a maximum value (~0.32 mmol/g) and the hyster-
esis, are very small. As shown in Fig. 4(d), the hysteresis at 328 K
was negligible even though a very small positive hysteresis was
still observed. This is taken as an indication that the basal spacing
of illite remained constant and did not expand upon exposure to
high pressure CO
2
although its layered structure is very similar
to that of montmorillonite in Table 1. Therefore, it is not easy for
Fig. 4. Comparison of excess sorbed amount on clay minerals plotted versus pressure: (a) montmorillonite at 318 K, (b) montmorillonite at 328 K, (c) illite at 318 K, (d) illite
at 328 K, (e) sepiolite at 318 K, (f) sepiolite at 328 K.
P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715 709
high pressure CO
2
to penetrate the basal spacing of illite and the
sorption capacity of illite is very small.
Compared to montmorillonite and illite, sepiolite, which has a
relatively high surface area (309.56 m
2
/g), showed a high excess
sorption capacity. The maximum values at 318 and 328 K were
~3.11 and ~2.98 mmol/g (Fig. 4(e) and (f)). Like montmorillonite,
the excess sorbed amount increased gradually with pressure. The
amount then decreased near the critical pressure at 318 K, but
the decrease of the excess sorption isotherm at 328 K was shifted
to a pressure greater than critical pressure. The hysteresis was rel-
atively smaller than it was in montmorillonite and it was negligible
at the supercritical region. However, the decrease of the excess
sorption isotherm with pressure in the supercritical region was
more signicant at 318 than at 328 K. Two isotherms crossed at
pressures over 100 bar. Since it is a non-swelling mineral [17], such
phenomena may result from a different mechanism of montmoril-
lonite, which will be analyzed in a later section.
In this study, the density of CO
2
was calculated from Duans EOS
[34] and showed good agreement with the result from the REF-
PROP program (Version 8.0) by National Institute of Standards
and Technology (NIST EOS) [35] (see the Appendix). The same
value as the critical density was obtained at 97.5 bar for 318 K
and 115 bar for 328 K.
Fig. 5(a) and (b) show the excess sorption isotherms of the clay
minerals as a function of CO
2
density. The isotherms of montmoril-
lonite and illite did not show any special phenomena near the
value of critical density. Alternatively, the crossover point of the
sorption isotherms at two temperatures was observed near the
value of critical density only in sepiolite. The excess desorption iso-
therms of clay minerals showed a similar shape and pattern as the
excess sorption isotherms. However, the crossover point of the iso-
therms at two temperatures was observed in the supercritical con-
dition for only montmorillonite.
In CO
2
sorption on anthracite coals, the phase density can pro-
vide more valuable information on the fundamental quantity than
pressure or temperature because the maximum excess sorption
and signicant change of the excess isotherms occurred at a near
critical value [15,16,36,37]. However, in the CO
2
sorption of clay
minerals, the phase transition from the subcritical gaseous phase
to the supercritical uid by pressure is more important than the
density to evaluate the excess sorption capacity.
The excess sorption capacity can benet the comparison of the
sorption capacity among clay minerals at the same pressure condi-
tion. Alternatively, the absolute sorbed amount is more convenient
to estimate the sorption storage capacity of the storage site in
terms of the nanoscale interface. Many studies reported the
isotherm models of the supercritical sorption for various gases
and sorbents [21,3840]. The modied DubininRadushkevich
(DR) equation strongly predicts the supercritical sorption on
coals [39]. Recently, the following supercritical sorption model
(Eq. (10)) with a uniformly sorbed density in the pore was sug-
gested [40],
m
E
= q
A

1
Z
PM
RT

v
P
H
A
(10)
where m
E
is the excess sorbed amount, q
A
is the density of the
sorbed phase, Z is the compressibility factor, v
P
is the pore volume,
and H
A
is a fractional lling term.
By combining the modied DR equation with the above
sorption model (Eq. (10)), the absolute sorbed amount could be
estimated from the excess sorbed amount (see the Appendix).
Figs. 68 compare the experimental data and predicted excess iso-
therms using the model equation, indicating the experimental
excess sorbed amount could be well predicted by the model. And
the absolute isotherms are also presented in the gures, which
were calculated by Eq. (C5) (See Appendix). Since the critical den-
sity of CO
2
is 467.6 kg/m
3
, the pressure at the same value as critical
density was different at each temperature as shown in Figs. 68.
Under subcritical conditions, the absolute sorbed amount grad-
ually increased with pressure and the difference in the sorbed
amount between two temperature conditions increased with pres-
sure in all clay minerals. However, from higher value than the crit-
ical density of CO
2
, the increase in the absolute sorption amount
was very small (montmorillonite and sepiolite) and even slightly
decreased (illite) with an increase in pressure. Such unusual behav-
ior of illite can be derived from the reduction in surface area. Fur-
ther discussion of this phenomenon will present in the BET surface
area comparison.
The ideal material for determining the sorbed phase density is
activated carbon. The sorbed phase density of CO
2
for activated
carbon was 1 g/cm
3
[16,33]. However, the density of the sorbed
phase may be different for each clay mineral (Table 3). Illite shows
a very low sorbed density so that it results in the negative excess
sorption shown in Fig. 7. In this study, the values to calculate the
maximum sorption amount in Eq. (10) were obtained by tting
the excess sorbed isotherm with the modied DR equation and
the values of q
A
, v
P
, and D are listed in Table 3. The maximum
CO
2
sorption capacities of clay minerals, m
A
max
, were calculated
from the pore volume (v
P
) multiplied by the sorbed phase density
of (q
A
). The percentage of maximum CO
2
sorption capacities
(g CO
2
/g clay%) of montmorillonite and sepiolite were 165.0 and
Fig. 5. Comparison of excess sorbed amount on clay minerals plotted versus
density: (a) adsorption, (b) desorption.
710 P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715
446.6 wt%, respectively, while illite was only 43.16 wt% as shown
in Table 3.
3.2. Structural analysis of clay minerals
Structural change in clay minerals was carried out at many
studies [41,42]. In the study, the structural characteristics of clay
minerals were compared by N
2
sorption/desorption isotherms
before and after the exposure to CO
2
(Fig. 9). The N
2
isotherms at
77 K can be limited by activated diffusion into the internal
Fig. 6. Comparison of CO
2
sorbed amount on montmorillonite between experiment
and prediction.
Fig. 7. Comparison of CO
2
sorbed amount on illite between experiment and
prediction.
Table 3
Calculated model values for montmorillonite, illite, and sepiolite.
Montmorillonite Illite Sepiolite
318 K q
A
(g/cm
3
) 0.886 0.352 0.856
v
p
(cm
3
/g) 1.862 1.227 5.216
D 0.0887 0.0917 0.0555
m
A
max
(g CO
2
/g clay) 1.650 0.4316 4.466
328 K q
A
(g/cm
3
) 0.800 0.228 1.203
v
p
(cm
3
/g) 1.612 1.432 3.346
D 0.0863 0.1607 0.0458
m
A
max
(g CO
2
/g clay) 1.290 0.3263 4.025
Fig. 8. Comparison of CO
2
sorbed amount on sepiolite between experiment and
prediction.
Fig. 9. Comparison of N
2
isotherms at 77 K before and after experiments.
Fig. 10. Comparison of XRD patterns of montmorillonite before and after
experiments.
P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715 711
micropore structure [43]. However, the relative change of clay
minerals before and after high pressure CO
2
exposure can be inves-
tigated. The surface area, pore area, and pore size before and after
CO
2
experiments are summarized in Table 4. The average values of
illite at 318 and 328 K are presented due to their relatively small
values.
The pore structure of the clay minerals can be irreversibly chan-
ged by pore fracture and pore blocking at high pressure CO
2
. As
shown in Fig. 9, the N
2
isotherms of clay minerals after experi-
ments were lower than those of the original clay minerals. It is
reported that montmorillonite can swell reversibly upon 50 bar
[42]. In the Figure, montmorillonite exposed to high pressure CO
2
showed a similar hysteresis shape as the original montmorillonite.
However, the relative hysteresis pressure range of illite was chan-
ged. And hysteresis was observed for the sepiolite after the CO
2
sorption experiment, which was not found in the raw material
(see Fig. 9).
After CO
2
sorption experiments at high pressure, the BET sur-
face area of all clay minerals had decreased signicantly as shown
in Table 4, which is opposite to anthracite coal that exhibited an
increased surface area after exposure to high pressure CO
2
[16].
The swelling clay mineral, montmorillonite, showed a less signi-
cant decrease of surface area than the non-swelling clay minerals,
sepiolite and illite (approximately 50%), implying that, like coal,
the swelling of clay may contribute to less surface area decrease.
The micropore surface area of illite and sepiolite signicantly
decreased while montmorillonite showed a signicant decrease
of the mesopore surface area. An increase of the mesopore size
was observed in sepiolite, but the change in the pore size was neg-
ligible for montmorillonite and illite. Fig. 10 shows the XRD result
of montmorillonite before and after the experiments. The XRD
peaks of montmorillonite were identied as a combination of Na-
montmorillonite and Ca-montmorillonite (JCPDS No. 29-1499 and
13-0135, respectively). The shift of (001), (101), and (107) peaks
in montmonrillonite after the high pressure CO
2
sorption was not
observed that crystallinity didnt change in spite of reduction in
surface area. And the reduced intensity was observed in the basal
plane (001) even though montmorillonite is the swelling clay
mineral.
Although montmorillonite is known as swelling clay, the sur-
face area of mesopore in montmorillonite was signicantly
reduced after exposed to high CO
2
pressure. On the other hand,
the changes in the surface area of the micropore and the pore size
of mesopore were relatively small. It implies that the inside struc-
ture of the mesopore space was highly changed in montmorillon-
ite. Therefore, it was expected that the hysteresis of CO
2
sorption
isotherm in montmorillonite might stem from swelling as well
from structural change.
3.3. Kinetics of the CO
2
sorption on clay mineral
The uptake of CO
2
at both temperatures was observed up to
120 bar for the three clay minerals, and the representative kinetic
uptake curves of montmorillonite at 5 and 110 bar are presented in
Fig. 11.
Table 4
Comparison of BET and pore analysis before and after experiments.
N
2
isotherm analysis Montmorillonite Illite Sepiolite
Before After 318 K After 328 K Before After 318 K & 328 K Before After 318 K After 328 K
Surface area (m
2
/g) BET 87.77 65.18 63.55 23.25 12.72 309.6 143.7 162.7
Pore area (m
2
/g) Micropore DR 80.97 72.02 68.86 18.46 10.35 300.7 121.9 153.5
Mesopore BJH 73.49 23.67 28.29 30.56 20.81 172.2 157.8 167.0
Pore size (A) Mean mesopore BJH 37.81 42.98 34.06 34.30 38.47 36.76 68.22 85.05
Fig. 12. Comparison of apparent diffusion coefcients.
Fig. 11. Sorption uptake of CO
2
and the t of the Fickian diffusion model at 318 K on
montmorillonite: (a) 5 bar, (b) 110 bar.
712 P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715
The sorption uptake was measured at each pressure step until it
reached an equilibrium. As mentioned in the sorption measure-
ment section, 1600 min was applied to each experimental run at
rst. Then, the weight change of the clay mineral by CO
2
sorption
was continuously monitored to conrm the equilibrium until the
value was not changed in the range of 10
4
g/h. However, it took
a long time to attain the equilibrium and the increase of the sorbed
amount was minute in the last part of the uptake. Therefore, in this
study, the sorption rate was evaluated from the uptake curve dur-
ing a certain period of time, showing a steep change in the sorption
amount. The uptake data up to 1600 min (about over 90% of the
equilibrium sorption amount) was applied to the Fickian diffusion
model (Eq. (4)). Fig. 11 shows the experimental uptake of CO
2
on
the clay mineral and the predicted result of Fickian diffusion. Since
the variation of the sorption amount and uid density was very
sensitive in the supercritical range [44], data scattering was
observed as shown in Fig. 11(b).
The obtained apparent diffusion coefcients, D
app
/r
2
(s
1
), of
clay minerals are summarized in Fig. 12. The average value (sym-
bol) with error bar was evaluated from uptake curves at all exper-
imental pressure range. Montmorillonite showed slightly slower
diffusion at 328 K than at 318 K, and the diffusion in illite and sepi-
olite slightly increased with temperature. As shown in Table 4,
montmorillonite showed a signicant decrease in the surface area
of mesopore at both temperatures, compared to the other clay
minerals. In addition, after the sorption experiment, the mesopore
size increased a little at 328 K, but decreased a little at 328 K. How-
ever, since the sorption rate depended on the pore structure, tem-
perature and sorption afnity in Eqs. (5)(9), it was hard to discuss
quantitative comparison among the clay minerals. However,
regardless of the difference in surface area and sorption amount,
the sorption rates of the tested clay minerals were nearly the same
order of magnitude (10
8
m
2
/s), implying a very slow uptake of
CO
2
compared to other porous sorbents.
The structure and diffusion of pure CO
2
in slit pores between
clay plates was reported by molecular simulation studies [14,45].
The calculated self-diffusion coefcient of muscovite, an illite-like
mica as a proxy system of clay minerals, showed an order of mag-
nitude comparable to the experimental results in this study. The
effect of the pressure on diffusion is relatively small, which can
be attributed to small changes in the average sorbed density [14].
4. Conclusions
Clay minerals are one of the major constituents of rocks such as
cap rock, reservoir rock, and coal mineral matter. CO
2
sorption at
the nanoscale interfaces of montmorillonite, illite, and sepiolite
were studied from subcritical to supercritical conditions. The sorp-
tion isotherms and uptake of each clay mineral were measured by
a gravimetric method at 318 and 338 K up to 120 bar.
Sepiolite with the highest surface area showed a CO
2
excess
sorption isotherm greater than the other clay minerals over all
pressure ranges. The sorbed amount of CO
2
on illite, which is the
major constituent of shale, was the smallest. The maximum CO
2
sorption capacities of sepiolite, montmorillonite, and illite were
446.6, 165.0 and 43.16 wt%, respectively. In the CO
2
sorption/
desorption isotherms, montmorillonite showed a signicant
positive hysteresis, but the hysteresis on illite and sepiolite was
relatively weak. The excess sorption isotherms of all clay minerals
showed a maximum near the critical pressure. Therefore, the
excess sorption isotherm decreased after the CO
2
phase transition
from the subcritical to supercritical condition. Alternatively, the
absolute sorption isotherms of all clay minerals increased with
pressure, but their variations were negligible from higher value
than the critical density of CO
2
. The sorption rates of clay minerals
were within a single order of magnitude (10
8
m
2
/s). Therefore, the
sorption amount and clay structure were not important factors in
the sorption rate.
The BET surface area of all clay minerals showed a drastic
decrease after high pressure CO
2
sorption experiments. The
decrease in surface area of the swelling clay mineral, montmoril-
lonite, was less signicant than the non-swelling clay minerals,
sepiolite and illite (approximately 50%). However, the pore struc-
ture of clay minerals was irreversibly deformed due to exposure
to high pressure CO
2
.
According to the results, mineralogy of sedimentary rock is very
important in CO
2
geological storage because the change of property
and sorption capacity is different among clay minerals at high
pressure CO
2
conditions. In the study, the sorption mechanism
and behaviors of CO
2
at nanoscale interfaces of clay minerals can
contribute to evaluating the sorption capacity and sealing integrity
of sedimentary rock in CO
2
geological storage. An understanding of
the effect of water molecules on pore deformation and CO
2
capac-
ity is needed for further investigations of geological sequestration
of CO
2
.
Acknowledgments
The authors gratefully acknowledge the nancial support pro-
vided for this work by the Korea Carbon Capture & Sequestration
R&D Center grant funded by the Ministry of Science, ICT and Future
Planning of the Korean Government (no. 20120008929).
Appendix A. Method for sorption measurement
After placing the prepared sorbent sample of weight (m
sample
0
)
in the basket, the magnetic suspension balance was evacuated
and the weight (M
0
) under vacuum was measured.
M
0
= m
metal
m
0
sample
(A1)
where m
metal
represents the weight of the lifted metal parts.
The system was then lled with helium and the volume of the
metal parts and the sorbent sample, V
metal
V
0
sorbent
, was calculated
from the measured weight, M(q
b
He
; T), at density q
b
He
and tempera-
ture T. This was based on the assumption that helium was not
sorbed.
V
0
= V
metal
V
0
sorbent
=
M
0
M(q
b
He
; T)
q
b
He
(A2)
After evacuating the cell again, it was lled with sorbing gas and
the weight M(q
b
; T) was measured at the desired conditions, i.e.
gas density q
b
and temperature T.
M(q
b
; T) = M
0
m
sorbed
q
b
[V
metal
V
0
sorbent
V
sorbed
[ (A3)
where m
sorbed
and V
sorbed
are the absolute sorbed amount and volume
of the sorbed phase, respectively. Since the latter could not be
directly measured, the sorption amount was commonly represented
by excess sorption:
m
ex
(q
b
; T) = m
sorbed
q
b
V
sorbed
=M(q
b
; T) M
0
q
b
[V
metal
V
0
sorbent
[
(A4)
As shown in Eq. (A4), the calculation of excess sorption did
require information regarding the volume of the sorbed phase.
The right-hand side of the equation contained only measurable
variables, i.e. balance signal (M(q
b
; T)) at equilibrium, M
0
under
vacuum, gas density (q
b
), volume of the suspended metal parts
(V
metal
), and volume of the sorbent sample in the balance (V
0
sorbent
).
P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715 713
Appendix B. Sorption mathematical model
A model to calculate the total amount of CO
2
within the system
is described, taking into account the density variation of uid
within pores. This allows for an sorbate (sorbed phase) with a den-
sity, q
A
, greater than that of the bulk gas, q
B
[40]. The sorbate in the
pore can be categorized by three factors: excess sorbed amount on
the surface (m
E
), amount of CO
2
without interaction with the sur-
face (m
B(A)
), and absolute (total) amount of CO
2
within the sorbate
(m
A
). The combination of these factors represents the total amount
of sorbate within the pores, including the amount of bulk CO
2
, m
P
.
Excess sorbed amount is the total amount of sorbate within the
pore minus the bulk amount of sorbate within the pore. Since
excess sorbed amount is the only value measured with equipment,
the total sorbed amount within the pores including the amount of
bulk CO
2
in the pores can be presented as follows:
m
E
= m
A
m
B(A)
(B1)
where
m
A
= q
A
v
A
(B2)
where q
A
is the density of the sorbate, and v
A
is the volume of the
sorbate, and
m
B(A)
= q
B
v
A
(B3)
Therefore, substituting Eqs. (B2) and (B3) into Eq. (B1) gives
m
E
= q
A
v
A
q
B
v
A
(B4)
which can be simplied to give
m
E
= (q
A
q
B
)v
A
(B5)
Note that the total amount in the pore is
m
P
= m
E
q
B
v
P
(B6)
The bulk sorptive density can be determined using an equation
of state
q
B
=
1
Z
PM
RT
v
P
H
A
(B7)
where M is the molar mass and Z is the compressibility factor (for
an ideal gas, P ?0, Z ?1). Duans EOS was used to describe the
CO
2
behavior. A fractional lling term, H
A
, is dened as,
H
A
=
v
A
v
P
(B8)
H
A
is the form of an IUPAC Type I isotherm. Therefore, by
substituting Eqs. (B7) and (B8) into Eq. (B5), the following equation
can be derived
m
E
= q
A

1
Z
PM
RT

v
P
H
A
(B9)
Eq. (B9) is the model framework for analyzing experimental
excess sorption isotherms. Note that at a low pressure m
E
= m
A
= q
A
H
A
v
P
, which can lead to difculties in obtaining separate, sta-
ble estimates of the product q
A
v
P
from data tting.
Appendix C. Modied DubininRadushkevich equation
The classic DR sorption isotherm is widely used for determining
the sorption capacity of materials.
W
ads
= W
0
exp D ln
P
s
P

2
" #
(C1)
W
0
is the surface sorption capacity of the substrate, D is a con-
stant related to the afnity of the sorbent for the gas, and P
s
is the
saturation pressure (equal to the pressure at which the gas con-
denses at the temperature of the test). D and W
0
are the free
parameters whose values are determined by model tting. The
DR sorption isotherm is a pore-lling model; it does not assume
monolayer surface coverage. However, with certain assumptions
regarding the distribution of pore sizes, the surface areas can be
estimated from W
0
.
This form of the DR equation is only valid when the pressure is
less than the saturation pressure of the gas (P < P
s
). It also cannot
be used to model the sorption of gases at temperatures or pres-
sures where the gases are supercritical and P
s
is undened. Never-
theless, because it is so useful, attempts have been made to extend
the equations to supercritical temperatures, primarily by attempt-
ing to dene an equivalent saturation pressure at temperatures
above the supercritical limits. However, these adaptations cannot
easily accommodate sorption at conditions where both the pres-
sure and temperature are above the critical values. A detailed
investigation of the meaning of the DR isotherm [39,46,47] has
indicated that the term P
s
is not necessarily the saturation pres-
sure, but is related to the energy required to compress the gas to
the sorbed phase density at the sorbent surface. If so, then there
is no physical reason why a form of the DR isotherm cannot be
applied in supercritical conditions, since, even in supercritical con-
ditions, the density of the sorbed gas is greater than the free gas
density.
It was recently proposed that the DR isotherm can be applied to
supercritical conditions by replacing the P
s
term with the sorbed
phase density, q
A
, and pressure with bulk gas density, q
B
[39]. This
modied DR isotherm can be used over a much wider pressure and
temperature range than available. This modied DR isotherm takes
the form of
W
sorbed
= W
0
exp D ln
q
A
q
b

2
" #
(C2)
In this paper, Eq. (C2) was converted to fractional lling H
A
Eq. (C3),
H
A
=
W
sorbed
W
0
= exp D ln
q
A
q
b

2
" #
(C3)
and combined with Eq. (B9) to yield Eq. (C4)
m
E
= q
A

1
Z
PM
RT

v
P
exp D ln
q
A
q
b

2
" #
(C4)
Fig. S1. CO
2
density comparison between NIST and Duans EOS.
714 P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715
The experiment value was tted to Eq. (C4) to get q
A
, v
P
and D.
The absolute sorbed amount was calculated using Eq. (C5)
m
A
= q
A
v
A
= q
A
v
P
H
A
= q
A
v
P
exp D ln
q
A
q
b

2
" #
(C5)
Appendix D. Density comparison between NIST EOS and Duans
EOS
CO
2
density from NIST EOS and Duans EOS were compared, and
there was little difference Fig. S1.
References
[1] T. Fujii, Y. Sato, H. Lin, H. Inomata, T. Hashida, Evaluation of CO
2
sorption
capacity of rocks using a gravimetric method for CO
2
geological sequestration,
Energy Procedia 1 (2009) 37233730.
[2] S. Bachu, W.D. Gunter, E.H. Perkins, Aquifer disposal of CO
2
: hydrodynamic and
mineral trapping, Energy Convers. Manage. 35 (1994) 269279.
[3] S. Bachu, Sequestration of CO
2
in geological media: criteria and approach for
site selection in response to climate change, Energy Convers. Manage. 41
(2000) 953970.
[4] S. Holloway, Safety of the underground disposal of carbon dioxide, Energy
Convers. Manage. 38 (1997) S241S245.
[5] A. Hildenbrand, S. Schlmer, B.M. Krooss, R. Littke, Gas breakthrough
experiments on pelitic rocks: comparative study with N
2
, CO
2
and CH
4
,
Geouids 4 (2004) 6180.
[6] J.Z.T.M. Al-Basali, M.M. Sharma, in: Annual Technology Conference and
Exhibition, SPE, Houston, Texas, 2005.
[7] J. Song, D. Zhang, Comprehensive review of caprock-sealing mechanisms for
geologic carbon sequestration, Environ. Sci. Technol. 47 (2012) 922.
[8] A. Busch, S. Alles, Y. Gensterblum, D. Prinz, D.N. Dewhurst, M.D. Raven, H.
Stanjek, B.M. Krooss, Carbon dioxide storage potential of shales, Int. J.
Greenhouse Gas Control 2 (2008) 297308.
[9] A. Busch, S. Alles, B.M. Krooss, H. Stanjek, D. Dewhurst, Effects of physical
sorption and chemical reactions of CO
2
in shaly caprocks, Energy Procedia 1
(2009) 32293235.
[10] D.J. Garcia, H. Shao, Y. Hu, J.R. Ray, Y.-S. Jun, Supercritical CO
2
-brine induced
dissolution, swelling, and secondary mineral formation on phlogopite surfaces
at 7595 C and 75 atm, Energy Environ. Sci. 5 (2012) 57585767.
[11] H. Shao, J.R. Ray, Y.-S. Jun, Dissolution and precipitation of clay minerals under
geologic CO
2
sequestration conditions: CO
2
brinephlogopite interactions,
Environ. Sci. Technol. 44 (2010) 59996005.
[12] W.D. Gunter, S. Wong, D.B. Cheel, G. Sjostrom, Large CO
2
sinks: their role in the
mitigation of greenhouse gases from an international, national (Canadian) and
provincial (Alberta) perspective, Appl. Energy 61 (1998) 209227.
[13] S. Bachu, J.J. Adams, Sequestration of CO
2
in geological media in response to
climate change: capacity of deep saline aquifers to sequester CO
2
in solution,
Energy Convers. Manage. 44 (2003) 31513175.
[14] D.R. Cole, A.A. Chialvo, G. Rother, L. Vlcek, P.T. Cummings, Supercritical uid
behavior at nanoscale interfaces: Implications for CO
2
sequestration in
geologic formations, Philos. Mag. 90 (2010) 23392363.
[15] H.J. Kim, Y. Shi, J. He, H.-H. Lee, C.-H. Lee, Adsorption characteristics of CO
2
and
CH
4
on dry and wet coal from subcritical to supercritical conditions, Chem.
Eng. J. 171 (2011) 4553.
[16] J. He, Y. Shi, S. Ahn, J.W. Kang, C.-H. Lee, Adsorption and desorption of CO
2
on
Korean coal under subcritical to supercritical conditions, J. Phys. Chem. B 114
(2010) 48544861.
[17] F.A. Mumpton, R. Roy, New data on sepiolite and attapulgite, Clays Clay Miner.
5 (1958) 136143.
[18] R.W. Mooney, A.G. Keenan, L.A. Wood, Adsorption of water vapor by
montmorillonite. I. Heat of desorption and application of BET theory 1, J.
Am. Chem. Soc. 74 (1952) 13671371.
[19] J. Thomas Jr., B.F. Bohor, Surface area of montmorillonite from the dynamic
sorption of nitrogen and carbon dioxide, Clays Clay Miner. 16 (1968) 8391.
[20] L. Aylmore, I. Sills, J. Quirk, Surface area of homoionic illite and
montmorillonite clay minerals as measured by the sorption of nitrogen and
carbon dioxide, Clays Clay Miner. 18 (1970) 9196.
[21] Y. Zhou, L. Zhou, Fundamentals of high pressure adsorption, Langmuir 25
(2009) 1346113466.
[22] F. Dreisbach, H.W. Lsch, P. Harting, Highest pressure adsorption equilibria
data: measurement with magnetic suspension balance and analysis with a
new adsorbent/adsorbate-volume, Adsorption 8 (2002) 95109.
[23] S. Ottiger, R. Pini, G. Storti, M. Mazzotti, Competitive adsorption equilibria of
CO
2
and CH
4
on a dry coal, Adsorption 14 (2008) 539556.
[24] S.W. Rutherford, J.E. Coons, Equilibrium and kinetics of water adsorption in
carbon molecular sieve: theory and experiment, Langmuir 20 (2004) 8681
8687.
[25] H.K. Chagger, F.E. Ndaji, M.L. Sykes, K.M. Thomas, Kinetics of adsorption and
diffusional characteristics of carbon molecular sieves, Carbon 33 (1995) 1405
1411.
[26] N.J. Foley, K.M. Thomas, P.L. Forshaw, D. Stanton, P.R. Norman, Kinetics of
water vapor adsorption on activated carbon, Langmuir 13 (1997) 20832089.
[27] D. Do Duong, Adsorption Analysis: Equilibria and Kinetics, Imperial College Pr,
1998.
[28] D.M. Ruthven, Principles of Adsorption and Adsorption Processes, John Wiley &
Sons, 1984.
[29] V.N. Romanov, A.L. Goodman, J.W. Larsen, Errors in Co
2
adsorption
measurements caused by coal swelling, Energy Fuels 20 (2005) 415416.
[30] R. Sakurovs, S. Day, S. Weir, Causes and consequences of errors in determining
sorption capacity of coals for carbon dioxide at high pressure, Int. J. Coal Geol.
77 (2009) 1622.
[31] T.J. Tambach, P.G. Bolhuis, B. Smit, A molecular mechanism of hysteresis in
clay swelling, Angew. Chem. 116 (2004) 27042706.
[32] R. Yong, Swelling pressures of sodium montmorillonite at depressed
temperatures, Clays Clay Miner. 11 (1962) 268281.
[33] M. Sudibandriyo, Z. Pan, J.E. Fitzgerald, R.L. Robinson, K.A.M. Gasem,
Adsorption of methane, nitrogen, carbon dioxide, and their binary mixtures
on dry activated carbon at 318.2 K and pressures up to 13.6 MPa, Langmuir 19
(2003) 53235331.
[34] Z. Duan, N. Mller, J.H. Weare, An equation of state for the CH
4
CO
2
H
2
O
system: I. Pure systems from 0 to 1000 C and 0 to 8000 bar, Geochim.
Cosmochim. Acta 56 (1992) 26052617.
[35] E.W. Lemmon, M.L. Huber, M.O. McLinden, NIST Reference Fluid
Thermodynamic and Transport Properties-REFPROP, vol. 80305, Physical and
Chemical Properties Division, National Institute of Standards and Technology,
Boulder, Colorado, 2007.
[36] H.-H. Lee, H.-J. Kim, Y. Shi, D. Keffer, C.-H. Lee, Competitive adsorption of CO
2
/
CH
4
mixture on dry and wet coal from subcritical to supercritical conditions,
Chem. Eng. J. 230 (2013) 93101.
[37] Z. Majewska, G. Ceglarska-Stefan ska, S. Majewski, J. Zie tek, Binary gas
sorption/desorption experiments on a bituminous coal: simultaneous
measurements on sorption kinetics, volumetric strain and acoustic emission,
Int. J. Coal Geol. 77 (2009) 90102.
[38] K. Kaneko, K. Murata, An analytical method of micropore lling of a
supercritical gas, Adsorption 3 (1997) 197208.
[39] R. Sakurovs, S. Day, S. Weir, G. Duffy, Application of a modied
DubininRadushkevich equation to adsorption of gases by coals under
supercritical conditions, Energy Fuels 21 (2007) 992997.
[40] J. Sharpe, N. Bimbo, V. Ting, A. Burrows, D. Jiang, T. Mays, Supercritical
hydrogen adsorption in nanostructured solids with hydrogen density variation
in pores, Adsorption 19 (2013) 643652.
[41] V.N. Romanov, Evidence of irreversible CO
2
intercalation in montmorillonite,
Int. J. Greenhouse Gas Control 14 (2013) 220226.
[42] P. Giesting, S. Guggenheim, A.F. Koster van Groos, A. Busch, Interaction of
carbon dioxide with Na-exchanged montmorillonite at pressures to 640 bars:
implications for CO
2
sequestration, Int. J. Greenhouse Gas Control 8 (2012) 73
81.
[43] O.P. Mahajan, CO
2
surface area of coals: the 25-year paradox, Carbon 29 (1991)
735742.
[44] K. Nishikawa, A.A. Arai, T. Morita, Density uctuation of supercritical uids
obtained from small-angle X-ray scattering experiment and thermodynamic
calculation, J. Supercrit. Fluids 30 (2004) 249257.
[45] X. Yang, C. Zhang, Structure and diffusion behavior of dense carbon dioxide
uid in clay-like slit pores by molecular dynamics simulation, Chem. Phys.
Lett. 407 (2005) 427432.
[46] G.F. Cerofolini, The DubininRadushkevich (DR) equation: history of a problem
and perspectives for a theory, in: D.H. Everett (Ed.), Colloid Science, The Royal
Society of Chemistry, 1983, pp. 5983.
[47] G.F. Cerofolini, Multilayer adsorption on heterogeneous surfaces, J. Low Temp.
Phys. 6 (1972) 473486.
P.R. Jeon et al. / Chemical Engineering Journal 255 (2014) 705715 715

Vous aimerez peut-être aussi