You are on page 1of 8

ARTICLE

pubs.acs.org/IECR

Suitability of a Solid Amine Sorbent for CO2 Capture by Pressure Swing


Adsorption
A. D. Ebner,† M. L. Gray,‡ N. G. Chisholm,† Q. T. Black,† D. D. Mumford,† M. A. Nicholson,† and
J. A. Ritter*,†

Department of Chemical Engineering, University of South Carolina, Columbia, South Carolina 29208, United States

National Energy Technology Laboratory, U.S. Department of Energy, P.O. Box 10940, Pittsburgh, Pennsylvania 15236, United States

ABSTRACT: This study showed that a solid amine, composed of poly(ethylenimine) immobilized into a CARiACT G10 silica
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

substrate, is a feasible sorbent for applications in a pressure swing adsorption (PSA) process for postcombustion CO2 capture. This
deduction materialized from an extensive study of the behavior of this material over a wide range of industrially relevant conditions
using thermogravimetric analysis. The temperature ranged from 40 to 100 °C, the CO2 partial pressure ranged from 1.2 to 100 vol %
with the total pressure fixed at 1 atm, the relative humidity ranged from dry conditions to 2 vol %, and the number of consecutive
adsorption and desorption cycles ranged from 4 to 76. The results revealed that this solid amine sorbent was very stable under the
conditions investigated. Water vapor at a low relative humidity exhibited only a minor and reversible effect on both the
Downloaded via CSIRO on January 19, 2023 at 03:06:39 (UTC).

thermodynamics and kinetics of the CO2 uptake and release. The isothermal CO2 working capacity ranged between 0.25 and 2.8
mol/kg, increased with increasing CO2 concentration, exhibited a maximum with increasing temperature, and produced a heat of
adsorption/reaction of around 50.0 kJ/mol. It was also determined that the optimal operating temperature for a PSA process was
around 80 °C for CO2 partial pressures >10 kPa and 6070 °C for CO2 partial pressures <10 kPa.

’ INTRODUCTION solid amine sorbent,1446 a relatively new type of CO2 sorbent, has
A considerable effort is underway worldwide to reduce the been reported to have these desirable characteristics.
amount of CO2 emissions from coal-fired and other fossil-fuel- Solid-supported amine sorbents are comprised of a high-
based power plants because these plants are responsible for over surface-area support, such as Amberlite,15,16 poly(methyl metha-
40% of the emissions of greenhouse gas in the USA alone.1 The crylate),1719 and polystyrene,19 but mostly silica,1845 with amine
goal is to capture CO2 from a stack or flue gas (i.e., postcombus- functional groups immobilized or grafted to its surface. The amine
tion capture), concentrate it to over 95 vol %, and sequester it within these sorbents behaves similarly to a liquid alkanolamine and
somewhere in the Earth. Viable CO2 capture options include thus takes up appreciable amounts of CO2 reversibly, even in, or
absorption, cryogenic condensation, adsorption, and membrane improved by, the presence of water vapor.19,2126 They are being
technologies.1 To date, however, none of these technologies has studied extensively not only for natural gas purification27 and pre-
been economically attractive; so, considerable research is being and postcombustion CO2 capture,1526,2845 i.e., bulk CO2 re-
done on all of them in search of a more economically feasible moval, but also for trace CO2 removal from air in spacecraft cabins
process. In particular, pressure swing adsorption (PSA) offers and submarines.1517 Although solid amine sorbents are being
features that make it attractive over the other processes. These touted for bulk CO2 capture using a temperature swing adsorption
features include its ease of retrofitting, the requirement of (TSA) process and although they are the first choice by NASA for
nothing more than electricity from the facility, and accomplish- trace CO2 removal from air by PSA, they have not been considered
ing regeneration without the need of facility and equipment- in detail for bulk CO2 removal from stack or flue gas by PSA, except
intensive heat or steam. in the recent work by Belmabkhout and Sayari.41
To date, hydrotalcites18 and zeolites1,913 are respectively Therefore, the objective of this study was to further determine
being studied for high- (precombustion) and low-temperature if a solid amine sorbent is suitable for the capture of bulk CO2 by
(postcombustion) capture of CO2. Hydrotalcites have one of the PSA. To this end, a solid amine sorbent, consisting of poly(ethyl-
highest CO2 working capacities in the 300500 °C range,1 enimine) (PEI) supported on a commercial silica substrate, was
whereas zeolites have one of the highest CO2 working capacities investigated over a wide range of industrially relevant conditions.
in the 3080 °C range.1 However, a major disadvantage of The parameters of interest included temperature, CO2 partial
zeolites for postcombustion CO2 capture is the need to remove pressure, water vapor content, and cycling stability. The kinetic
water vapor from the feed stream because of its strong affinity and thermodynamic information generated from this study is
with zeolites, which impairs CO2 capture. suitable for correlation with a mechanistic model, which can then
Flue gas typically contains up to 1015 vol % H2O, which
must be removed by using either a guard bed prior to the PSA unit Received: January 12, 2011
or a layered adsorbent bed in the PSA unit. Both approaches add Accepted: March 23, 2011
complexity and cost. Hence, a CO2 sorbent that is water-tolerant Revised: March 6, 2011
and exhibits a sufficient CO2 working capacity is highly desirable. A Published: April 02, 2011

r 2011 American Chemical Society 5634 dx.doi.org/10.1021/ie2000709 | Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

syringe pump filled with water and placed in the feed line. The
speed of the syringe pump was such that it provided a feed gas
containing about 2.0 vol % water vapor continuously during both
the adsorption and desorption steps. To ensure immediate and
complete evaporation of the liquid water exiting the needle of the
syringe, a heating band and thermocouple were placed at the port
connecting the needle to the feed line to generate a temperature
of about 200 °C at the tip of the needle. Four cycles were carried
out with the syringe pump turned on, and then the same sample
was activated again and four more cycles were carried out with
the syringe pump turned off (i.e., under dry conditions). Two
sets of wet conditions were used: 80 °C with a feed containing 1.2
vol % CO2 (dry basis) and 40 °C with a feed containing 100 vol %
CO2 (dry basis).

’ EQUILIBRIUM MODEL DEVELOPMENT


Figure 1. Schematic of the adsorptiondesorption cycling experimen- The mechanism envisioned to describe the kinetics of the CO2
tal apparatus, depicting a thermogravimetric analyzer, an electronic valve chemisorption reaction with the CARiACT G10 solid amine
timer, feed gases, and a water syringe pump.
sorbent was defined in the simplest terms according to the
following mass balance:
be used in a dynamic adsorption process simulator to study the
performance of a PSA process. DqCO2
¼ kf qCO2 , a ðN  qCO2 Þ  kb qCO2 ð1Þ
Dt
’ MATERIAL PREPARATION AND CYCLE TESTING where qCO2 represents the chemisorbed CO2 that has reacted
Reagent-grade methanol and PEI (MN 423) were purchased with the immobilized amine, N represents the total number of
from Aldrich Chemical Co. and used as received in the prepara- sites available for this reaction, qCO2,a represents unreacted CO2
tion of the solid amine sorbent. The substrate for the solid amine in a physisorbed state and within the vicinity of some of the
sorbent was CARiACT G10 silica from Fuji Silysia. It has a immobilized amine, and kf and kb represent kinetic constants for
surface area of 300 m2/g, a pore volume of 1.2 mL/g, and particle the forward and backward reactions.
sizes between 200 and 400 μm. The amine was dissolved in This analysis also assumed that the forward reaction depends
methanol and combined with silica at an amine/methanol/silica on qCO2,a. In other words, the immobilized amine reacts directly
weight ratio of 0.67:2.0:1.0. This slurry was placed in a rotary with physisorbed CO2 and not gaseous CO2. If qCO2,a is assumed
evaporator to remove methanol and allow amine to physically to be small and in equilibrium with the partial pressure of CO2
adsorb (i.e., become immobilized) in the silica beads. The according to Henry’s law, i.e.,
resulting solid amine sorbent contained about 40 wt % PEI.
qCO2 , a ¼ kH PCO2 ð2Þ
Further details on this synthesis method, including the use of
other solid amine sorbents, are provided elsewhere.19 then at equilibrium (i.e., ∂qCO2/∂t = 0) the chemisorbed CO2 can
A Perkin-Elmer TGA-7 thermogravimetric analyzer was used be defined in terms of a Langmuir-type expression according to
to measure the dynamic adsorption and desorption behavior of
CO2 on the solid amine sorbent. A schematic of the experimental kPCO2 N
qCO2 ¼ ð3Þ
apparatus is displayed in Figure 1. First, a sample (∼20 mg) was 1 þ kPCO2
activated at 100 °C for 80 min in N2 (UHP grade, Airgas) flowing
at around 60 cm3/min and 1 atm. At the end of the activation with
step, the temperature was adjusted to a predetermined value  
kf kH ΔH
between 40 and 100 °C using a 20 °C/min ramping rate. When k¼ ¼ k0 exp  ð4Þ
the temperature was reached, the test gas was switched from N2 kb RT
to a N2/CO2 gas mixture (also flowing at around 60 cm3/min and
and 1 atm) to initiate adsorption and begin the first half of an
adsorptiondesorption cycle. The concentration of CO2 ΔH ¼ Ef þ ΔHa  Eb ð5Þ
(Coleman grade, Airgas) in the mixture was varied between 1
and 100 vol %. The adsorption step was continued for 40 min, k is the affinity coefficient between CO2 and the sorbent, ΔH is
and then the gas was switched back to N2 to initiate desorption the effective heat of adsorption/reaction, Ef and Eb are the
and finish the second half of the adsorptiondesorption cycle. activation energies for the forward and backward reactions, and
Unless otherwise indicated, four adsorptiondesorption cycles ΔHa is the heat of adsorption associated with physisorbed CO2.
were carried to elucidate the CO2 loading and dynamic behavior To obtain ΔH and k0 from experimental data, eq 3 can be
of the amine-doped CARiACT G10. To study the stability of the expressed in the linear form
material, 78 cycles were carried out continuously with a sample at PCO2 1 1
80 °C using 100 vol % CO2 during the adsorption step. ¼ þ PCO2 ð6Þ
qCO2 kN N
The role of humidity on the adsorptiondesorption behavior
of CO2 on the CARiACT G10 solid amine sorbent was evaluated where 1/kN and 1/N can be resolved respectively from the y
in a similar fashion but with the aid of a Cole Parmer 74900 series intercept and slope of a plot of PCO2/qCO2 versus PCO2.
5635 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

Figure 2. CO2 loading in a CARiACT G10 solid amine sorbent at 80 °C


for 78 adsorption (CO2) and desorption (N2) cycles with a 40 min half-
cycle time. The gas flow rate was set at 60 cm3/min and 1 atm. Cycles
1861 were removed for clarity.

’ RESULTS AND DISCUSSION


The stability of the CARiACT G10 solid amine sorbent is
displayed in Figure 2. It shows the adsorption and desorption
behavior during 78 continuous cycles. Cycles 1861 were Figure 3. CO2 loading in a CARiACT G10 solid amine sorbent at (a)
removed for clarity. This run was carried out at 80 °C under 80 °C for four adsorption (CO2 at 1.2 vol % in N2, dry basis) and
dry conditions using 100 vol % CO2 during the adsorption step. desorption (N2, dry basis) cycles with a 40 min half-cycle time and (b)
The performance of this sorbent did not show any signs of 40 °C for four adsorption (CO2, dry basis) and desorption (N2, dry
deterioration during the 78 cycles, as noted by there being no basis) cycles with a 40 min half-cycle time while continuously exposing
apparent difference between the behavior of cycle 17 and cycle 62 the sample to 2 vol % H2O vapor in the flowing gas stream, the same
sample under dry conditions after reactivation, and a completely
and the working capacity (i.e., the difference between the CO2 different sample under dry conditions. The gas flow rate was set at
loading at the end of the adsorption step and the CO2 loading at 60 cm3/min and 1 atm.
the end of the desorption step) being constant at around 2.75
mol/kg (∼12 wt %). The stability of this particular solid amine
sorbent is thus suitable for use in a PSA process. shows a run carried out at 40 °C using a feed containing 100 vol %
As an aside, it is worth noting that this result is in stark contrast CO2 (dry basis) during the adsorption step.
to that recently reported by Sayari’s group.21 They showed a The presence of H2O vapor had a slight effect on both the
significant decrease in the working capacity while cycling a kinetics and thermodynamics associated with the uptake and
similar PEI-based solid amine sorbent at 105 °C under dry release of CO2 by the sorbent. This was particularly true when
conditions. They attributed the loss of capacity to the possible the relative humidity was higher at 40 °C. However, the perfor-
formation of urea. However, Drage et al.46 reported that urea mance of the sorbent returned to that exhibited under dry
formation takes place only at temperatures above 135 °C, which conditions when the sorbent was subsequently reactivated and
is well above the temperatures run in this work and that by run under dry conditions. Moreover, the curves for the two
Sayari’s group.21 The stability of the CARiACT G10 solid amine different samples of sorbent run under dry conditions nearly
sorbent under dry conditions might be due to the presence of a overlapped, suggesting that the effect of H2O vapor on the
trace amount of water (∼50 ppm) that leaked in during the behavior of the sorbent was completely reversible, as reported
thermogravimetric analysis (TGA) runs that protected the elsewhere with other solid amine sorbents.19,2126 This material
sample or due to differences in the preparation procedures. is thus far superior to zeolites with respect to processing a water-
The effect of water vapor on the adsorption and desorption saturated feed stream because it obviates the use of a guard bed or
behavior of CO2 on the CARiACT G10 solid amine sorbent a layered bed in a PSA process.
under low humidity conditions is displayed in Figure 3. It shows The effects of the temperature and CO2 concentration on the
four continuous cycles that were operated under humid condi- adsorption and desorption behavior of CO2 on the CARiACT
tions during both the adsorption and desorption steps by G10 solid amine sorbent are shown in Figures 4 and 5. They each
injecting 2 vol % H2O vapor into the flowing gas stream. The display four cycles at 40, 60, 80, and 100 °C and at different CO2
corresponding relative humidities at 80 and 40 °C are 4.27% and concentrations (in N2) during the adsorption step of 1.2, 4.8,
27.45%, respectively. For comparison, the figure also includes 14.5, and 32.8 vol % (Figure 4) and 56.1, 69.8, 88.6, and 100 vol
four cycles carried out under identical but dry conditions on the % (Figure 5). The panels on the left show the behavior of the
same sorbent after it was reactivated and four cycles carried out sorbent throughout all four cycles, while the panels on the right
under dry conditions on a completely different sorbent. Figure 3a show the behavior of the samples during the first 4 min of the
shows a run carried out at 80 °C using a feed containing 1.2 vol % adsorption step of the first cycle. All runs were carried under dry
CO2 (dry basis) in N2 during the adsorption step, while Figure 3b conditions.
5636 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

Figure 4. CO2 loading in a CARiACT G10 solid amine sorbent at 40, 60, 80, and 100 °C for four adsorption [CO2 at (a) 1.2, (b) 4.8, (c) 14.5, and (d)
32.8 vol % in N2] and desorption (N2) cycles with a 40 min half-cycle time during (1) all four cycles and (2) the first 4 min of the adsorption step during
the first cycle. The gas flow rate was set at 60 cm3/min and 1 atm.

This solid amine sorbent exhibited significant, but widely vary- In contrast, the very slow desorption rates at 40 °C limited the
ing, working capacities that ranged between 0.25 (Figure 4a.1, working capacity to about 0.40.6 mol/kg for all CO2 partial
100 °C) and 2.8 mol/kg (Figure 5d.1, 80 °C) for a broad range of pressures. The thermodynamic limitation arises because the
temperatures and CO2 partial pressures. There were two plausible reaction between CO2 and the amine is exothermic, which
reasons why the working capacity of this material was limited under decreased the chemisorbed CO2 conversion with increasing tem-
certain conditions. One reason was associated with kinetics, and the perature. Notice that the maximum CO2 loadings achieved by the
other reason was associated with thermodynamics. samples were consistently smaller at 100 °C than at 80 °C, with the
The kinetic limitation arises because the CO2 desorption effect being significantly more pronounced at lower CO2 partial
kinetics were strongly dependent on and thus limited by the pressures (Figure 4a.1). This compromise between the kinetic
temperature. Only the runs carried out at 80 and 100 °C achieved limitation at lower temperatures and the thermodynamic limitation
complete regeneration during the 40 min desorption step in N2. at higher temperatures revealed a range of temperatures between
5637 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

Figure 5. CO2 loading in a CARiACT G10 solid amine sorbent at 40, 60, 80, and 100 °C for four adsorption [CO2 at (a) 56.1, (b) 69.8, (c) 88.6, and (d)
100.0 vol % in N2] and desorption (N2) cycles with a 40 min half-cycle time during (1) all four cycles and (2) the first 4 min of the adsorption step during
the first cycle. The gas flow rate was set at 60 cm3/min and 1 atm.

40 and 100 °C that exhibit an optimal CO2 working capacity for a These figures also show that the rate of CO2 adsorption
PSA process. increased with decreasing temperature, giving rise to an apparent
Another interesting aspect shown in Figures 4 and 5 was the negative activation energy. This intriguing result can be explained
relatively fast initial rate of CO2 adsorption compared to that by the presence of physisorbed CO2 that not only was in
observed during desorption, which was then followed by a very thermodynamic equilibrium with CO2 in the gas phase but also
slow uptake of CO2 that was more apparent in the runs carried out at participated directly in the reaction with the immobilized amine,
40 and 60 °C. Except for the 1.2 vol % CO2 case, all of the runs as indicated by the mass balance expression in eq 1. According to
exhibited this very fast dynamic in the first 5 min of the adsorption eq 2, which represents the equilibrium state of the physisorbed
step (panels on the right). These characteristically different rates of CO2, the rate of the forward reaction in eq 1 becomes controlled
adsorption alluded to the presence of a multistep adsorption/ by an activation energy given by Ef þ ΔHa that may take on
reaction mechanism taking place during CO2 uptake. negative values. Because Ef is always a positive number and ΔHa
5638 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

Figure 7. Model predictions versus experimental data: (a) equilibrium


CO2 loading in a CARiACT G10 solid amine sorbent at 80 and 100 °C
according to eq 6 (linearized version of eq 3); R2 = 0.9989; (b)
equilibrium CO2 loading in a CARiACT G10 solid amine sorbent at
80 and 100 °C according to eq 3 and equilibrium CO2 loading predicted
from eq 3 at 40 and 60 °C by extrapolation. All experimental CO2
loadings correspond to those from the last (fourth) cycle shown in
Figures 4 and 5. An apparent equilibrium state is reached in the CO2
loading at 80 and 100 °C but not at 40 and 60 °C (refer to the text).

physisorbed CO2 in equilibrium with PCO2 that participated


directly in the reaction with the immobilized amine. In contrast,
the CO2 desorption rates were much smaller, strongly decreased
with decreasing temperature, and exhibited only a weak depen-
dence on PCO2, particularly at the lower temperatures, i.e., at 40
and 60 °C. This apparent dependence of the desorption rate on
PCO2 was actually manifested through its dependence on the
initial CO2 loading at the onset of the desorption step. As a result
of the thermodynamic restriction described earlier, those initial
Figure 6. (a) Adsorption rates of CO2 in a CARiACT G10 solid amine CO2 loadings generally decreased with increasing temperature
sorbent at different temperatures and CO2 partial pressures during the and decreasing PCO2, giving rise to this apparent dependence of
adsorption step of the first cycle. (b) Desorption rates of CO2 in the
the desorption rate on PCO2.
amine-doped CARiACT G10 at different temperatures and CO2 partial
pressures during the desorption step of the first cycle. (c) CO2 working It was mentioned earlier that it would be possible to operate a
capacity in a CARiACT G10 solid amine sorbent at different tempera- PSA process in the 40100 °C range, based on the CO2 working
tures and CO2 partial pressures during the last (fourth) cycle. capacities observed in Figures 4 and 5. Figure 6c further shows that
the optimal operating temperature for a PSA process depended on
is always a negative number, it is entirely possible that the PCO2. For situations where the CO2 partial pressure in the feed is
absolute magnitude of ΔHa can be greater than that of Ef, giving larger than about 10 kPa, around 80 °C would be optimal. However,
rise to this apparent negative activation energy.47 for situations where the CO2 partial pressure in the feed is below
A clearer depiction of the trend between the adsorption rate about 5 kPa, between 60 and 80 °C would be optimal.
and temperature is shown in Figure 6. It displays the maximum It was also very clear from the runs carried out at 80 and
adsorption (Figure 6a) and desorption (Figure 6b) rates during 100 °C (Figures 4 and 5) that an equilibrium state was achieved
the first cycle of the runs shown in Figures 4 and 5 and the CO2 by the end of the adsorption step at all of the CO2 partial
working capacity of the last (fourth) cycle of those same runs pressures. It was also very clear from the runs carried out at 40
(Figure 6c). In addition to a strong dependence of the CO2 and 60 °C that this was not the case at those lower temperatures
adsorption rate on PCO2, the CO2 adsorption rates were generally because of the very slow kinetic process that took place after the
faster at lower temperatures. This was again consistent with initial rapid uptake of CO2 in the first 5 min. Figure 7 shows the
5639 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641
Industrial & Engineering Chemistry Research ARTICLE

correlation of eq 3 or 6 to the experimental equilibrium CO2 far superior to zeolites. An optimal PSA operating temperature
loadings extracted from the runs carried out at just 80 and was determined to be around 80 °C for CO2 partial pressures >10
100 °C. Figure 7a displays the very good fit (R2 = 0.9989) that kPa and between 60 and 70 °C for CO2 partial pressures <10 kPa.
resulted by fitting simultaneously all of those loadings to the These temperature ranges are ideally suited for postcombustion
linearized form of eq 3, i.e., eq 6, over the entire range of CO2 CO2 capture.
partial pressures. The resulting linear behavior suggested that the
underlying thermodynamic mechanism described by eq 1 was ’ AUTHOR INFORMATION
quite plausible.
Corresponding Author
The corresponding fitting parameters from eq 6 were N =
*E-mail: ritter@cec.sc.edu.
2.891 mol/kg, k0 = 1.069  108 kPa1, and ΔH = 5.003  101
kJ/mol. These values were consistent with those previously
reported for this material.19 They were also used in eq 3 to ’ ACKNOWLEDGMENT
generate the curves in Figure 7b, which displays a comparison of
the experimental data with the Langmuir-type model at 80 and The authors gratefully acknowledge financial support pro-
100 °C. The agreement was again quite good. However, when vided, in part, by the NSF Research Experience for Under-
eq 3 was used to extrapolate the equilibrium behavior to the two graduates (REU) Site in Chemical Engineering at the
lower temperatures (40 and 60 °C), the predictions did not agree University of South Carolina under Grant EEC-0851997 and,
with the experimental CO2 loadings extracted from the corre- in part, by the Center for Strategic Approaches to the Generation
sponding runs at 40 and 60 °C at the end of the adsorption step of of Electricity at the University of South Carolina.
the fourth cycle (Figures 4 and 5). This was not surprising and
further confirmed that equilibrium was not attained at those ’ NOMENCLATURE
lower temperatures because of the very slow kinetic process that Eb = backward activation energy for the CO2 chemisorption
was taking place. These results suggested very strongly that the reaction (J/mol)
optimal temperature for operating a PSA process would be in the Ef = forward activation energy for the CO2 chemisorption
80 °C range, where both the kinetics (adsorption and desorption reaction (J/mol)
rates) and thermodynamics (CO2 loading) would be most k = affinity coefficient between CO2 and the sorbent (1/kPa)
favorable. k0 = preexponential constant for the affinity coefficient (1/kPa)
kb = backward rate constant for the CO2 chemisorption reaction (1/s)
’ CONCLUSIONS kf = forward rate constant for the CO2 chemisorption reaction
(kg/mol/s)
A solid amine sorbent consisting of PEI immobilized into a kH = Henry’s law constant for physisorbed CO2 (kg/mol/kPa)
CARiACT G10 silica substrate was investigated over a wide N = total number of sites for chemisorbed CO2 (mol of CO2/kg of
range of conditions using TGA to evaluate its potential for CO2 sorbent)
capture by PSA. Each TGA run consisted of an initial activation PCO2 = partial pressure of CO2 in the gas phase (kPa)
step at 100 °C for 80 min in N2, followed by four consecutive qCO2 = loading of chemisorbed CO2 (mol of CO2/kg of sorbent)
cycles of 40 min adsorption in CO2/N2 and 40 min desorption in qCO2,a = loading of physisorbed CO2 (mol of CO2/kg of sorbent)
N2. The role of the CO2 concentration in the adsorption step R = universal gas constant (J/mol/K)
(1.2, 4.8, 14.5, 32.8, 56.1, 69.8, 88.6, and 100 vol %) at four T = temperature (K)
temperatures (40, 60, 80, and 100 °C) under dry conditions was
investigated, along with the role of water vapor at two conditions Greek Symbols
(2 vol % H2O and 1.2 vol % CO2 at 80 °C and 2 vol % H2O and ΔHa = heat of adsorption for physisorbed CO2 (J/mol)
100 vol % CO2 at 40 °C). The sorbent stability was also ΔH = effective heat of adsorption/reaction (J/mol)
investigated by carrying out 78 consecutive adsorption (100 vol %
CO2) and desorption (100 vol % N2) cycles at 80 °C under dry ’ REFERENCES
conditions.
The results revealed that CARiACT G10 solid amine sorbent (1) Ebner, A. D.; Ritter, J. A. State-of-the-art Adsorption and
Membrane Separation Processes for Carbon Dioxide Production from
was very stable over 78 adsorptiondesorption cycles under the
Carbon Dioxide Emitting Industries. Sep. Sci. Technol. 2009, 44, 1273.
conditions investigated. Water vapor at a low relative humidity (2) Reijers, T. H.; Valster-Schiermeier, S. E.; Cobden, P. D.; Brink.,
exhibited only a minor and reversible effect on both the thermo- R. W. Hydrotalcite as CO2 Sorbent for Sorption-Enhanced Steam
dynamics and kinetics of CO2 uptake and release. The isothermal Reforming of Methane. Ind. Eng. Chem. Res. 2006, 45, 2522.
CO2 working capacity ranged between 0.25 and 2.8 mol/kg, (3) Ding, Y.; Alpay, E. Equilibria and Kinetics of CO2 Adsorption on
increased with increasing CO2 concentration, exhibited a max- Hydrotalcite Adsorbent. Chem. Eng. Sci. 2000, 55, 3461.
imum with increasing temperature, and produced a heat of (4) Lee, K. B.; Beaver, M. G.; Caram, H. S.; Sircar, S. Reversible
adsorption/reaction of around 50.0 kJ/mol. Chemisorption of Carbon Dioxide: Simultaneous Production of Fuel-
Overall, these results suggested that CARiACT G10 solid cell grade H2 and Compressed CO2 from Synthesis Gas. Adsorption
amine sorbent is suitable for use in a PSA process for post- 2007, 13, 385.
(5) Oliveira, E. L.; Grande, C. A.; Rodrigues, A. E. CO2 Sorption on
combustion CO2 capture, based on its attractive kinetic and
Hydrotalcite and Alkali-modified (K and Cs) Hydrotalcite at High
thermodynamic properties. Because it was very stable with Temperatures. Sep. Purif. Technol. 2008, 62, 137.
respect to isothermal cycling and in the presence of water vapor (6) Walspurger, S.; Cobden, P. D.; Safonova, O. V.; Wu, Y. H.;
and because it exhibited reasonable CO2 uptake and release rates Anthony, E. J. High CO2 Storage Capacity in Alkali-Promoted Hydro-
and working capacities over a broad range of temperatures, the talcite-Based Material: In Situ Detection of Reversible Formation of
performance of this material in a PSA process should prove to be Magnesium Carbonate. Chem.—Eur. J. 2010, 16, 12694.

5640 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641


Industrial & Engineering Chemistry Research ARTICLE

(7) Du, H.; Ebner, A. D.; Ritter, J. A. Temperature dependence of (28) Tsuda, T.; Fujiwara, T.; Taketani, Y.; Saegusa, T. Amino silica-
Non-Equilibrium Kinetic Model that Describes the Reversible Adsorp- gels acting as a carbon-dioxide absorbent. Chem. Lett. 1992, 11, 2161.
tion and Desorption Behavior of CO2 in a K-Promoted HTlc. Ind. Eng. (29) Tsuda, T.; Fujiwara, T. Polyethyleneimine and macrocyclic
Chem. Res. 2010, 49, 3328. polyamine silica-gels acting as carbon-dioxide absorbents. J. Chem. Soc.,
(8) Du, H.; Ebner, A. D.; Ritter, J. A. Pressure dependence of Non- Chem. Commun. 1992, 22, 1659.
Equilibrium Kinetic Model that Describes the Reversible Adsorption (30) Li, W.; Bollini, P.; Didas, S. A.; Drese, J. H.; Choi, S.; Jones,
and Desorption Behavior of CO2 in a K-Promoted HTlc. Ind. Eng. Chem. C. W. Structural Changes of Silica Mesocellular Foam Supported
Res. 2011, 50, 412. Amine-Functionalized CO2 Adsorbents Upon Exposure to Steam.
(9) Agarwal, A.; Biegler, L. T.; Zitney, S. E. A Superstructure-Based ACS Appl. Mater. Interfaces 2010, 2, 3363.
Optimal Synthesis of PSA Cycles for Post-Combustion CO2 Capture. (31) Li, W.; Choi, S.; Drese, J. H.; Hornbostel, M.; Krishnan, G.;
AIChE J. 2010, 56, 1813. Eisenberger, P. M.; Jones, C. W. Steam-Stripping for Regeneration of
(10) Mosca, A.; Hedlund, J.; Webley, P. A.; Grahn, M.; Rezaei, F. Supported Amine-Based CO2 Adsorbents. Chem. Sustainable Chem.
Structured zeolite NaX coatings on ceramic cordierite monolith sup- 2010, 3, 899.
ports for PSA applications. Microporous Mesoporous Mater. 2010, (32) Hicks, J. C.; Drese, J. H.; Fauth, D. J.; Gray, M. L.; Qi, G. G.;
130, 38. Jones, C. W. Designing adsorbents for CO2 capture from flue gas-
(11) Tlili, N.; Grevillot, G.; Vallieres, C. Carbon dioxide capture and hyperbranched aminosilicas capable of capturing CO2 reversibly. J. Am.
recovery by means of TSA and/or VSA. Int. J. Greenhouse Gas Control Chem. Soc. 2008, 130, 2902.
2009, 3, 519. (33) Lee, S.; Filburn, T. P.; Gray, M. L.; Park, J. W.; Song, H. J.
(12) Zhang, J.; Singh, R.; Webley, P. A. Alkali and alkaline-earth Screening test of solid amine sorbents for CO2 capture. Ind. Eng. Chem.
cation exchanged chabazite zeolites for adsorption based CO2 capture. Res. 2008, 19, 7419.
Microporous Mesoporous Mater. 2008, 111, 478. (34) Khatri, R. A.; Chuang, S. S. C.; Soong, Y.; Gray, M. L. Thermal
(13) Zhang, J.; Webley, P. A; Xiao, P. Effect of process parameters on and chemical stability of regenerable solid amine sorbent for CO2
power requirements of vacuum swing adsorption technology for CO2 capture. Energy Fuels 2006, 20, 1514.
capture from flue gas. Energy Convers. Manage. 2008, 49, 346. (35) Gray, M. L.; Soong, Y.; Champagne, K. J.; Pennline, H.; Baltrus,
(14) Choi, S.; Drese, J. H.; Jones, C. W. Adsorbent Materials for J. P.; Stevens, R. W.; Khatri, R.; Chuang, S. S. C.; Filburn, T. Improved
Carbon Dioxide Capture from Large Anthropogenic Point Sources. immobilized carbon dioxide capture sorbents. Fuel Process. Technol.
Chem. Sustainable Chem. 2009, 2, 796. 2005, 86, 1449.
(15) Birbara P. J.; Filburn, T. P.; Nalette, T. A. Regenerable solid (36) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Modeling CO2
amine sorbent. U.S. Patent 5,876,488, 1999. adsorption on amine-functionalized mesoporous silica: 1. A semi-
(16) Birbara P. J.; Nalette T. A. Regenerable supported ami- empirical equilibrium model. Chem. Eng. J. 2010, 161, 173.
nepolyol sorbent. U.S. Patent 5,492,683, 1996. (37) Serna-Guerrero, R.; Sayari, A. Modeling adsorption of CO2 on
(17) Satyapal, S.; Filburn, T.; Trela, J.; Strange, J. Performance and amine-functionalized mesoporous silica. 2: Kinetics and breakthrough
properties of a solid amine sorbent for carbon dioxide removal in space curves. Chem. Eng. J. 2010, 161, 182.
life support applications. Energy Fuels 2001, 15, 250. (38) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Influence of
(18) Gray, M. L.; Champagne, K. J.; Fauth, D.; Baltrus, J. P.; Pennline, regeneration conditions on the cyclic performance of amine-grafted
H. Performance of immobilized tertiary amine solid sorbents for the mesoporous silica for CO2 capture: An experimental and statistical
capture of carbon dioxide. Int. J. Greenhouse Gas Control 2008, 2, 3. study. Chem. Eng. Sci. 2010, 65, 4166.
(19) Gray, M. L.; Hoffman, J. S.; Hreha, D. C.; Fauth, D. J.; Hedges, (39) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Amine-bearing
S. W.; Champagne, K. J.; Pennline, H. W. Parametric Study of Solid mesoporous silica for CO2 removal from dry and humid air. Chem. Eng.
Amine Sorbents for the Capture of Carbon Dioxide. Energy Fuels 2009, Sci. 2010, 65, 3695.
23, 4840. (40) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Adsorption of
(20) Belmabkhout, Y.; Serna-Guerrero, R.; Sayari, A. Adsorption of CO2 from dry gases on MCM-41 silica at ambient temperature and high
CO2-Containing Gas Mixtures over Amine-Bearing Pore-Expanded pressure. 1: Pure CO2 adsorption. Chem. Eng. Sci. 2009, 64, 3721.
MCM-41 Silica: Application for Gas Purification. Ind. Eng. Chem. Res. (41) Belmabkhout, Y.; Sayari, A. Isothermal versus Non-isothermal
2010, 49, 359. Adsorption-Desorption Cycling of Triamine-Grafted Pore-Expanded
(21) Sayari, A.; Belmabkhout, Y. Stabilization of Amine-Containing MCM-41 Mesoporous Silica for CO2 Capture from Flue Gas. Energy
CO2 Adsorbents: Dramatic Effect of Water Vapor. J. Am. Chem. Soc. Fuels 2010, 24, 5273.
2010, 132, 6312. (42) Franchi, R.; Harlick, P. J. E.; Sayari, A. A high capacity, water
(22) Serna-Guerrero, R.; Belmabkhout, Y.; Sayari, A. Further in- tolerant adsorbent for CO2: diethanolamine supported on pore-ex-
vestigations of CO2 capture using triamine-grafted pore-expanded panded MCM-41. Nanoporous Mater. IV 2005, 156, 879.
mesoporous silica. Chem. Eng. J. 2010, 158, 513. (43) Franchi, R. S.; Harlick, P. J. E.; Sayari, A. Applications of pore-
(23) Drese, J. H.; Choi, S.; Lively, R. P.; Koros, W. J.; Fauth, D. J.; expanded mesoporous silica. 2. Development of a high-capacity, water-
Gray, M. L.; Jones, C. W. SynthesisStructure-Property Relationships tolerant adsorbent for CO2. Ind. Eng. Chem. Res. 2005, 44, 8007.
for Hyperbranched Aminosilica CO2 Adsorbents. Adv. Funct. Mater. (44) Hicks, J. C.; Dabestani, R.; Buchanan, A. C.; Jones, C. W.
2009, 19, 3821. Assessing site-isolation of amine groups on aminopropyl-functionalized
(24) Belmabkhout, Y.; Sayari, A. Effect of pore expansion and amine SBA-15 silica materials via spectroscopic and reactivity probes. Inorg.
functionalization of mesoporous silica on CO2 adsorption over a wide Chim. Acta 2008, 361, 3024.
range of conditions. Adsorption 2009, 15, 318. (45) Hicks, J. C.; Dabestani, R.; Buchanan, A. C.; Jones, C. W. Spacing
(25) Serna-Guerrero, R.; Da’na, E.; Sayari, A. New Insights into the and site isolation of amine groups in 3-aminopropyl-grafted silica materials:
Interactions of CO2 with Amine-Functionalized Silica. Ind. Eng. Chem. The role of protecting groups. Chem. Mater. 2006, 18, 5022.
Res. 2008, 47, 9406. (46) Drage, T. C.; Arenillas, A.; Smith, K. M.; Snape, C. E. Thermal
(26) Khatri, R. A.; Chuang, S. S. C.; Soong, Y.; Gray, M. L. Carbon stability of polyethylenimine based carbon dioxide adsorbents and its
dioxide capture by diamine-grafted SBA-15: A combined Fourier trans- influence on selection of regeneration strategies. Microporous Mesopor-
form infrared and mass spectrometry study. Ind. Eng. Chem. Res. 2005, ous Mater. 2008, 116, 504.
44, 3702. (47) Davis, M. E.; Davis, R. J. Fundamentals of Chemical Reaction
(27) Belmabkhout, Y.; De Weireld, G.; Sayari, A. Amine-Bearing Engineering; McGraw-Hill: New York, 2003.
Mesoporous Silica for CO2 and H2S Removal from Natural Gas and
Biogas. Langmuir 2009, 25, 13275.

5641 dx.doi.org/10.1021/ie2000709 |Ind. Eng. Chem. Res. 2011, 50, 5634–5641

You might also like