Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal Pre-proof

Insights into the CO2 capture over amine-functionalized mesoporous


silica adsorbents derived from rice husk ash

W. Henao, L.Y. Jaramillo, D. López, M. Romero-Sáez, R.


Buitrago-Sierra

PII: S2213-3437(20)30711-9
DOI: https://doi.org/10.1016/j.jece.2020.104362
Reference: JECE 104362

To appear in: Journal of Environmental Chemical Engineering

Received Date: 27 June 2020


Revised Date: 25 July 2020
Accepted Date: 30 July 2020

Please cite this article as: Henao W, Jaramillo LY, López D, Romero-Sáez M, Buitrago-Sierra
R, Insights into the CO2 capture over amine-functionalized mesoporous silica adsorbents
derived from rice husk ash, Journal of Environmental Chemical Engineering (2020),
doi: https://doi.org/10.1016/j.jece.2020.104362

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Insights into the CO2 capture over amine-functionalized mesoporous silica
adsorbents derived from rice husk ash

W. Henao a, L.Y. Jaramillo b, D. López c, M. Romero-Sáez d, R. Buitrago-Sierra a,*

a
Grupo Materiales Avanzados y Energía - MATyER, Facultad de Ingeniería, Instituto Tecnológico
Metropolitano, Calle 54A No. 30-01, Medellín, Colombia.
b
Grupo de Investigación en Tecnologías de Información y Medio Ambiente - GITIMA, Facultad de
Ingeniería, Tecnológico de Antioquia, Calle 78B No. 72A-220, Medellín, Colombia.
c
Química de Recursos Energéticos y Medio Ambiente, Instituto de Química, Facultad de Ciencias

of
Exactas y Naturales, Universidad de Antioquia UdeA, Calle 70 No. 52-21, Medellín, Colombia.
d
Grupo Química Básica, Aplicada y Ambiente - ALQUIMIA, Facultad de Ciencias Exactas y

ro
Aplicadas, Instituto Tecnológico Metropolitano, Calle 73 No. 76A-354, Medellín, Colombia

* Corresponding author: robinsonbuitrago@itm.edu.co -p


Graphical abstarct
re
lP
na
ur
Jo

1
Highlights

 RHA-derived silica adsorbents were effective for CO2 capture under mild conditions.
• Pore features of silica supports played a key role in the CO2 adsorption behavior.
• The importance of amine dispersion and efficient pore filling was highlighted.
• Kinetic and thermodynamic regimes dominated the CO2 adsorption at different levels.

ABSTRACT

The design of high-performance porous adsorbents active for CO2 capture is imperative to
mitigate the global climate problems arising from the accumulation of anthropogenic emissions. This

of
work evaluates the CO2 adsorption behavior of a series of amine-functionalized silicas with different
pore structures: SBA-15 (2D hexagonal), SBA-11 (3D cubic), and SiO2 (d) (disordered). The materials

ro
were synthesized using Rice Husk Ash (RHA) as a silica source and then functionalized with
polyethyleneimine (PEI) by wet impregnation. CO2 adsorption performance was found to be quite
-p
sensitive to the pore features of the silica supports and the impregnated amount of PEI. Among the
prepared adsorbents, the PEI/SBA-15 exhibited the highest amine utilization (0.38 mol CO2/mol N, at
20 wt.% PEI) and CO2 adsorption capacity (61.6 mg CO2/ g ads., at 40 wt.% PEI) under mild
re
conditions (40 ºC and ambient pressure). The outstanding performance of this adsorbent was attributed
to its uniform 2D hexagonal arrangement of cylindrical mesopores that decreases the CO2 mass
lP

transfer resistance and favors the PEI distribution through the pore network, enhancing the interaction
with the CO2 stream. Further evaluation of the adsorption kinetics indicated that the CO2 capture was
influenced by kinetic and thermodynamic regimes to different extents depending on the adsorption
na

temperature.

Keywords: CO2 capture, Rice husk ash, silica, SBA-15, SBA-11, polyethylenimine.
ur
Jo

1. INTRODUCTION

Increasing global emissions of CO2 is an urgent issue to be faced due to its detrimental
repercussions on a wide variety of environmental, economic and health aspects [1]. Despite the
implementation of alternative energy sources and strategies to reduce the anthropogenic CO2
emissions, the cumulative concentration of CO2 in the atmosphere is still an unavoidable concern. The
current challenge lies in the intimate dependence between the global energy demand and the burning
of fossil fuels and biomass, which constitute the major source of worldwide CO2 emissions [2].

2
Mitigation of the rising post-combustion CO2 levels while meeting the growing global energy demand
can only be achieved by implementing feasible materials and technologies that allow the carbon
capture and its subsequent utilization [3,4].

Scrubbing flue gases in chimneys exhaust with liquid alkanolamines is one of the most mature
technologies for CO2 capture in large-scale processes due to its high CO2 selectivity and facile
adaptation to existing power plants [5]. Nonetheless, some drawbacks such as low absorption capacity,
high equipment corrosion, fast solvent degradation and large energy consumption for regeneration,
hinder its performance under repeated CO2 adsorption-desorption cycles [6]. To overcome these
obstacles, the immobilization of amine-containing compounds over solid supports has attracted
considerable attention in the last decades [7]. Amine-modified solid adsorbents combine the high
affinity of amines towards CO2 along with the textural properties and thermal stability of porous

of
materials. Compared to liquid alkanolamines, solid adsorbents provide a more cost-effective cyclical
capture process and longer operational lifetime [8].

ro
The reaction between CO2 and solid amine-based adsorbents occurs via a zwitterion mechanism,
-p
where either carbamates or bicarbonates can be reversibly formed [9]. As shown in Eq. 1-2, under dry
conditions, 2 mol of primary or secondary amines react with 1 mol of CO2 to form carbamates. In
addition, in the presence of water, the reaction stoichiometry changes to 1:1 (CO2:NH) for the
re
formation of ammonium bicarbonates.

2R1R2NH + CO2 ↔ R1R2NH2+ + R1R2NCOO− (carbamate) (1)


lP

R1R2NH + CO2 + H2O ↔ R1R2NH2+ + HCO3− (bicarbonate) (2)

Several amine-containing molecules such as (3-Aminopropyl)triethoxysilane (APTES), (3-


na

Aminopropyl) trimethoxysilane (APTMS), tetraethylenepentamine (TEPA), pentaethylenehexamine


(PEHA), linear, branched and hyperbranched polyethylenimine (PEI) have been used to strengthen the
affinity of solid adsorbents towards the preferential adsorption of CO2 [10]. These amine groups can
ur

be incorporated into the porous material by in situ functionalization or post-synthesis through chemical
grafting or physical impregnation [11]. Chemical grafting with organosilanes provides low CO2
Jo

adsorption capacities because the number of structural amine groups supplied is quite low [12].
Conversely, physical impregnation with polymeric amines like PEI promotes a high CO2 adsorption
performance because the density of amine groups exposed to the flue gas is high. Nonetheless, the
thermal stability of the amine moieties is low due to weak interaction with the support, and the amount
of polymer to be incorporated must be carefully controlled in order to prevent CO2 diffusional
limitations through the porous network of the adsorbent [13].

3
Inorganic and organic porous materials including zeolites, silicas, metal-organic frameworks
(MOFs) and activated carbons have been extensively investigated as supports for amine
immobilization [14–18]. Among them, porous silicas are attractive supports because of their adjustable
pore features, large surface area, excellent chemical/thermal stability, and facile surface
functionalization. Commercial silica materials are usually obtained from tetraethyl orthosilicate
(TEOS) or sodium silicate (Na2SiO3) as silicon precursors [19]. However, TEOS is an expensive and
moisture-sensitive reagent, while Na2SiO3, often produced by smelting quartz sand and sodium
carbonate, requires a large amount of energy for its manufacturing increasing the silica cost [20]. In
this regard, the use of alternative raw materials to fabricate silica-based adsorbents still needs further
exploration to compensate for the economic and environmental shortcomings of their current
commercial counterparts.

of
Rice husk is well known as a highly siliceous biomass and an abundant agricultural waste derived
from the rice production chain. In 2019, the worldwide rice harvest was about 496.5 MMT of which

ro
20 wt.% corresponds to the husk [21]. Due to its high calorific value, rice husk is commonly used as a
fuel for energy production by direct combustion or gasification [22]. However, the resulting Rice Husk
-p
Ash (RHA) triggers environmental pollution as well as disposal problems, and therefore its post-
utilization is certainly necessary [23].
re
Amorphous silica can be easily extracted from RHA under alkaline conditions resulting in a
Na2SiO3 solution which is subsequently precipitated into SiO2 at pH values ranging from 3 to 10 [24].
lP

Furthermore, pore features such as size, volume and spatial arrangement can be finely adjusted during
precipitation by using surfactants as pore structure-directing templates through a cooperative self-
assembly approach. A large variety of silicas with different pore systems such as 2D-hexagonal (SBA-
na

15, MCM-41), 3D-bicontinuous cubic (MCM-48, KIT-6, SBA-16), 3D-cage-type cubic (SBA-11,
KIT-5) and lamellar (MCM-50) have been obtained through this methodology [25–27]. The pore
ur

structure parameters strongly influence the performance of the adsorbent for gas adsorption, in specific
towards CO2 capture, and therefore the comprehension of this interdependence is essential for
designing novel efficient adsorbents [28,29].
Jo

Hence, intensive research efforts have focused on understanding the role of parameters such as
pore size, pore volume and surface area of the silica supports, as well as of the features of the
incorporated amine functionalities on the CO2 capture performance of amine-supported adsorbents
[30–33]. R. Kishor and A. K. Ghoshal [30] evaluated a series of PEHA-impregnated mesoporous
silicas MCM-41, HV MCM-41 (high pore volume), KIT-6, and SBA-15 having different textural
properties. They found that, although specific surface area had little impact during CO2 adsorption

4
performance, large pore sizes provide easier access for CO2 to reach the amine active sites and prevent
the blockage of the channels, while large pore volumes can improve loading capacity of amine species
facilitating their dispersion and exposure to the CO2 stream. In this regard, multimodal porous
materials have attracted great attention because the advantage of their multiple pore sizes [34–36]. For
instance, J. Yu et al. [35] prepared bimodal meso-macroporous SiO2 hollow spheres achieving a high
CO2 adsorption capacity of 194 mg CO2/g ads at 110 °C when loaded with TEPA at 50 wt.%. The 3–
4 nm mesopores in the shell and the 103-117 nm macropores in the cavity of the hollow spheres led to
high dispersion of TEPA, and thus more CO2 affinity sites are exposed to the adsorbate. The influence
of the structure and molecular weight of supported PEI amino-polymer on the CO2 capture have also
been evaluated [37–39]. K. Li et al. [37] found that branched PEI promotes higher CO2 saturated
sorption capacities compared to linear PEI; however, it is less stable during CO2 adsorption–desorption

of
cycling. On the other hand, CO2 adsorption capacity decreased as PEI molecular weight increases,
while the CO2 regeneration heat was much lower than that of MEA solution for the supported PEI-

ro
silica adsorbents.

In the present work, the synergic effect of the pore features and the amine functionalization degree
-p
of a series of silicas with distinct pore systems (2D hexagonal, 3D cubic, disordered) on their CO2
adsorption performance is evaluated. The silica supports were prepared by using surfactants with
re
different molecular dimensions as pore structure-directing templates and RHA as a renewable silica
source. The resulting materials were amine-functionalized by PEI impregnation at different loadings,
lP

and their CO2 adsorption behavior was investigated at different operating temperatures. To further
understand the effect of the PEI loading and the adsorption temperature on the CO2 capture, a kinetic
study was conducted on the experimental data results.
na

2. EXPERIMENTAL
ur

2.1. Alkaline extraction of SiO2 from rice husk ash


Jo

Highly siliceous RHA (93 wt.% SiO2) obtained from rice husk combustion at 700 °C in a bubbling
fluidized bed reactor was used as a silica source [40]. 10 g of as-calcined RHA were refluxed with 500
mL of 2.0 M HCl at 110 °C for 4 h to leach the metallic impurities. Afterward, the treated RHA was
filtered and washed with copious amounts of boiling deionized water until the impurities were
removed. The washing was completed when the filtrated exhibited a constant electrolytic conductivity,
corroborated by using a multiparameter device (HI 9829, HANNA®). The recovered solid was dried
at 100 °C overnight to remove the moisture. Silica extraction was carried out by refluxing 5 g of treated

5
RHA with 50 mL of 2.5 M NaOH at 110 °C for 4 h under vigorous stirring (600-700 rpm). During
alkaline extraction, the SiO2 contained in the RHA reacts with NaOH yielding a sodium silicate
solution (Na2SiO3) and a carbonaceous residue. The resulting mixture was filtered retaining the carbon
residue in the filter while keeping the Na2SiO3 solution. The SiO2 content in this solution was
determined by gravimetric precipitation with 2.0 M HCl.

2.2. Synthesis of silica supports

Ordered mesoporous silicas were synthesized via cooperative self-assembly with Pluronic P123
(EO20PO70EO20, Sigma-Aldrich) or BrijC10 (C16H33(OCH2CH2)10OH, Sigma-Aldrich) as pore
structure-directing templates and Na2SiO3 (8.6 wt.% SiO2) from RHA as a silica source [41]. SBA-15
was obtained by dissolving 2 g P123 in 52.5 mL 2.0 M HCl subsequent dropwise addition of 20 mL

of
Na2SiO3 solution. The pH of the mixture was carefully adjusted to 3.0 and kept under stirring at 40 ºC
for 24 h. Later, the pH was adjusted to 7.0 to promote the SiO2 precipitation rate [42], maintaining the

ro
stirring for 1 h. The resultant gel was aged under static conditions at 100 ºC during 48 h.

SBA-11 silica with trimodal porosity was prepared by a dual templating approach based on a gel
-p
molar composition of 1 SiO2:0.010 P123:0.007 BrijC10: 5 HCl:161 H2O. First, the gel solutions of
P123 (G1) and BrijC10 (G2) were separately prepared. For G1 gel, P123 was dissolved in 2.0 M HCl
re
subsequent dropwise addition of Na2SiO3 under stirring for 3 h at 40 ºC and pH 3.0. G2 gel was
similarly prepared by stirring the BrijC10 in 2.0 M HCl + Na2SiO3 for 24 h at 40 ºC and pH 3.0. In a
lP

second step, the G1 gel was added to the G2 solution and the mixture was stirred for 1 h at 40 ºC and
pH 7.0. Later, the resultant mixed gel was aged under static conditions at 100 ºC during 48 h.

A third kind of silica with disordered pore structure was prepared without using porogen agents.
na

First, 20 mL Na2SiO3 solution were added dropwise to 52.5 mL 2.0 M HCl. Then, the pH of the mixture
was carefully adjusted to 3.0 and kept under stirring at 40 ºC for 24 h. The silica obtained through this
ur

method was labeled as SiO2 (d).

In all cases, the solid product was recovered by filtration and washed with copious amounts of
Jo

boiling deionized water to remove the NaCl by-product formed and/or the surfactant excess. The
washing was completed when the filtrated exhibited a constant electrolytic conductivity, corroborated
by using a multiparameter device (HI 9829, HANNA®). After drying at 110 ºC overnight, the solids
were ground in a mortar to obtain fine powders. The organic pore templates of SBA-15 and SBA-11
samples were removed by calcination in air at 550 ºC for 5 h with a heating rate of 1 °C/min.

2.3. PEI impregnation

6
All the derived-RHA silicas were impregnated with branched polyethyleneimine (b-PEI, Mn=
1200, ρ= 1.08 g/mL) using methanol as a solvent. The nominal content of PEI was adjusted to 20, 40
or 60 wt.% regarding the final amount of adsorbent. First, the appropriate amount of b-PEI was
dissolved in methanol and then the silica was added in a weight ratio of 1:32. The mixture was kept
under constant stirring at 50 ºC until evaporation of the solvent and a subsequent drying was performed
at room temperature overnight to avoid any amine degradation [43]. The impregnated materials were
designated as PEI(W)/S, where W and S represent the nominal amount of PEI (wt.%) and the silica
support used, respectively.

2.4. Materials characterization

Structural, morphological and porosity features of the prepared silicas were studied by low angle

of
X-ray Diffraction (XRD), Fourier Transform Infrared Spectroscopy (FTIR), Transmission Electron
Microscopy (TEM) and N2 physisorption. XRD patterns were recorded on a Philips PW-1830

ro
diffractometer using Cu Kα radiation (λ=1.5406 Å) from 0.5° to 3.5° 2θ degrees with a step size of
0.004°. FTIR spectra were collected in an IRTracer-100 spectrometer, from 400 to 4000 cm-1 with 15
-p
cumulative scans and a wavelength resolution of 4 cm-1. Powder samples were diluted in KBr with a
mass ratio of 1.0% and then pressed into disks for analysis. TEM analyses were conducted using a
re
Hitachi HT7700 microscope operated at 120 kV, previous dispersion of silicas by ultrasonic stirring
in isopropanol. N2 adsorption-desorption isotherms were measured at -196 ºC, using a Micromeritics
ASAP 2010 equipment. Prior to the analysis, samples were vacuum-degassed during 4 h at 350 °C for
lP

the silica supports and at 80 ºC during 1 h for the amino-containing materials. Surface area was
calculated by multipoint BET method (P/P0= 0.01-0.1), while total pore volume was obtained at the
maximum relative pressure reached by the adsorption branch (P/P0>0.99). Pore size distribution and
na

average pore diameter were obtained by the Barrett–Joyner–Halenda (BJH) method from the
desorption isotherm [44]. Micropores volume was estimated by the t-plot method using the Harkins-
ur

Jura thickness equation. PEI content of the silica adsorbents was determined in an SDT Q600 thermal
analyzer by heating the samples from 30 to 900 °C with a rate of 10 °C/min under oxidative atmosphere
Jo

(synthetic air 2.0 grade, 100 mL/min). Nitrogen content was determined in a LECO TruSpec Micro
elemental analyzer system.

2.5. CO2 adsorption tests

CO2 adsorption isotherms were measured under ambient pressure in an SDT Q600 thermal
analyzer (TA Instruments, USA). Approximately, 5 mg of adsorbent were first pretreated at 110 °C
for 1 h in N2 (50 mL/min) to remove any gas or moisture preadsorbed. Then the sample was cooled
down to 40 ºC maintaining the N2 flux (50 mL/min, 5.0 grade) and, once the temperature was reached,
7
a flow of 40 mL/min of pure CO2 (5.0 grade) was injected attaining a concentration of 44.4% CO2 /
55.6% N2 into the chamber. Sample mass evolution and temperature were continuously recorded
during 100 min of CO2 exposure. The adsorption capacity (mg CO2/g adsorbent) was determined as
the weight increase of the sample upon the CO2 adsorption experiment. The thermal stability of the
samples was evaluated by subjecting the best adsorbent to several CO2 adsorption-desorption cycles.
For this, the sample was regenerated in N2 atmosphere (50 mL/min) at 110 °C for 1 h after completion
of the first adsorption cycle, and then cooled down to 40 ºC to start the next cycle.

3. RESULTS AND DISCUSSION

3.1. Materials characterization

of
Pore structure information of the silica supports was obtained by x-ray diffraction, nitrogen
physisorption, and transmission electron microscopy. From the low-angle XRD patterns shown in Fig.
1A, it was observed that SBA-15 and SBA-11 samples present a sharp (100) diffraction peak at 0.9º

ro
2θ degrees, indicative of a well-ordered porous system. The other less intense Bragg reflections
detected in the 2θ range of 1.5º–3.0º provided insights into the spatial arrangement of the pores (Fig.
-p
1A, inset). In the case of SBA-15 silica, the three well-resolved peaks at 1.62º (110), 1.84º (200) and
2.36º (300) 2θ degrees are characteristic of a two-dimensional P6mm hexagonal pore system [45],
re
whereas the (210), (211) and (310) reflections centered at 1.90º, 2.06º and 2.55º in the SBA-11 pattern,
respectively, revealed the formation of a three-dimensional Pm3m cubic porous structure [46].
lP

Interestingly, the (100) and (110) peaks of SBA-15 were still present in the SBA-11 pattern, suggesting
that SBA-11 possesses to some extent the pore arrangement of SBA-15 silica as well. This is because
the Si-containing species condensed at the micelles periphery play as a capping agent that stabilizes
na

each individual micellar system, allowing their structural packing in well-defined mesostructures after
combination of the templates [41]. The low-angle XRD pattern of SiO2 (d) is not shown because it
does not exhibit any ordered pore system since no pore directing template was used during its
ur

synthesis.
Jo

8
Fig. 1. (A) Low-angle XRD patterns, (B) N2 adsorption-desorption isotherms and (C) pore size

of
distribution of synthesized silica supports.

ro
Fig. 1B shows the N2 adsorption-desorption isotherms of the synthesized silicas derived from
RHA. All the samples exhibited type IV isotherms arising from the N2 capillary condensation
occurring into the mesopores. Different hysteresis loops, namely type-H1 and type-H4, were observed
-p
suggesting the presence of pores with distinct shapes and sizes. SBA-15 exhibited a prominent type-
H1 hysteresis at 0.70< P/P0 <0.90, which is associated with uniform open-ended cylindrical mesopores
re
[47]. In the case of SiO2 (d), the capillary condensation step was narrower and occurred at lower
relative pressures 0.42< P/P0 <0.83 ascribed to smaller internal mesopores. Instead, SBA-11 revealed
lP

three hysteresis loops at 0.46< P/P0 <0.64 (type-H4), 0.64< P/P0 <0.84 (type-H1) and 0.84< P/P0 <0.99
(type-H4), suggesting the formation of a mixed mesopore system with three main pore sizes.
Accordingly, the pore size distributions depicted in Fig. 1C show that SBA-15 and SiO2 (d) silicas
na

present a narrow unimodal distribution with a mean pore size of 7.6 nm (FWHM= 1.5 nm) and 4.2 nm
(FWHM= 1.2 nm) respectively. On the other hand, SBA-11, prepared by a dual templating approach,
exhibited a trimodal pore distribution with mean diameters about 3.8 nm (FWHM= 0.7 nm), 7.7 nm
ur

(FWHM= 2.7 nm) and 12.9 nm (FWHM= 8.1 nm). This suggests that the micellar arrangement of each
surfactant template was preserved during the synthesis process due to the capping action of the silica
Jo

species condensed on its periphery [48].

Further information about the pore system of the silicas was obtained through the TEM images
displayed in Fig. 2. Different pore arrangements were observed depending on the pore-structure
directing template and the synthesis method performed. While SBA-15 presented a well-defined
hexagonal honeycomb structure of cylindrical mesopores, the SBA-11, prepared by a dual templating
approach, exhibited worm-like pores with a lower ordering degree. Contrariwise, SiO2 (d) did not show
any pore ordering as expected since no pore template was used during its synthesis, in agreement with
9
the XRD results. These differences in the pore structure of the silica supports make interesting the
understanding of its effect on the CO2 capture performance of the final developed adsorbents.

of
ro
-p
Fig. 2. Transmission electron micrographs of the synthesized silica supports.
re
FTIR spectra displayed in Fig. 3 allow the identification of the functional groups present in the
silicas before and after their impregnation with PEI. The bands centered at 470, 802 and 1097 cm-1
were attributed to the symmetrical stretching Si-O-Si (νs) and bending Si-O-Si (δ) vibrations of the
lP

silica supports, while the weak peak at 950 cm-1 was ascribed to the stretching vibration of the terminal
Si-OH [49]. The -OH vibration of the silica surface and the adsorbed water were identified at 3400
na

cm-1 and 650 cm-1, respectively [50]. After PEI impregnation, the appearance of new absorption bands
related to the vibrations of secondary amines (R2NH) at 1658 cm-1 and primary amines (RNH2) were
respectively observed at 1472 and 1590 cm-1 in all cases. In addition, the asymmetric and symmetric
ur

stretching modes of C-H were also identified at 2945 cm-1 at 2820 cm-1, which were attributed to the
chemical structure of PEI chains [51]. Naturally, the peak intensity of these bond vibrations was
Jo

increased with the rise of the PEI content in the final adsorbents.

10
of
ro
-p
re
lP

Fig. 3. FTIR spectra of the silicas before and after impregnation with PEI: (A) SBA-15, (B) SBA-11
and (C) SiO2 (d).

Thermogravimetric analysis of the adsorbents presented in Fig. S1 (Supplementary


na

Information) evidenced the advantageous thermal stability of the silica supports. Excluding the small
weight loss (~5 wt.%) about 100 ºC, attributed to moisture evaporation, no additional weight losses
ur

were identified from 100 °C to 900 ºC. This confirms that the pore-structure directing templates were
successfully removed, leaving the pores available for subsequent amine incorporation. Regarding the
PEI-impregnated adsorbents, a weight loss of ~10 wt.% was observed about 100 ºC ascribed to the
Jo

desorption of moisture and other pre-adsorbed gases [39]. Furthermore, a prominent weight loss was
detected from 200 ºC to 650 ºC, attributed to the thermal degradation and volatilization of the PEI
polymer [37]. This last weight loss was quite close to the nominal amount of PEI loading, which
confirms the effectiveness of the impregnation method. Since the thermal degradation of the PEI
occurs above 200 ºC, the activation and regeneration of the amine-modified silica adsorbents can be
carried out at 110 ºC without affecting the physicochemical properties of the PEI polymer.

11
The Textural properties of the bare silicas and the amine-supported adsorbents are summarized in
Table 1. SBA-15 silica exhibited both high BET surface area (604 m2/g) and total pore volume (1.192
cm3/g) originated from its rich porous structure of open-ended cylindrical pores with interconnected
wall micropores (t-plot micropore volume = 0.031 cm3/g). On the other hand, SBA-11 having trimodal
mesopores revealed a lower surface area (356 m2/g) and total pore volume (0.792 cm3/g), probably
because its worm-like pores limit the accessibility for the N2 probe molecules. In this case, the
micropore volume calculated by the t-plot method was 0.015 cm3/g. SiO2 (d) sample synthesized
without using a porogen agent showed the lowest surface area (256 m2/g) and pore volume (0.460
cm3/g).

After amine impregnation, both the surface area, pore volume and BJH pore size of the bare silica

of
supports were gradually decreased with the increasing of amine loading, confirming the filling of the
pore channels by the PEI polymer. As a result, the nitrogen sorption capacity was gradually decreased

ro
with the PEI content, which caused a modification of the hysteresis loops of the N2 isotherms and the
pore size distributions of the amine-modified silicas (Fig. S2-S4). During impregnation, PEI is
expected to first distribute into the mesoporous channels of the silica supports through a layer
-p
formation, coating the mesopores walls and occluding the micropores due to its large size. After
reaching some critical thickness, PEI begins to aggregate into a plug that grows in the axial direction
re
of the pore with further amine loading. Finally, when the occupied volume exceeds the pore volume
of the support, PEI polymer molecules spill out and coat the outer surface of the particles until a full
lP

coverage occurs [52,53]. Thus, the reduction in N2 adsorption indicates the extent of pore filling during
impregnation. The complete filling of the pores results in a restriction for further access to N2 gas, and
therefore, no reliable information on the textural properties of the adsorbents could be obtained at high
na

PEI loading contents.

Since each silica has a different porous structure capable to accommodate a specific amount of
ur

PEI, the efficient filling of pores is an important consideration for tuning the porosity of novel
adsorbents and the optimal amine loading as well. For instance, the pore volume saturation (99%
Jo

regarding the bare support) of the modified SBA-15 and SBA-11 silicas was attained at 60 wt.% PEI
loading, whereas the SiO2 (d) sample, exhibiting the lowest surface area and pore volume, reached the
saturation at 40 wt.% PEI loading. From the nitrogen isotherms and pore size distributions shown in
Fig. S2-S4, it was observed that the porous structure of the silica supports is preserved below this
critical value of PEI loading, suggesting that most of the polymer was distributed into the pore
channels, with a small fraction coating the outer surface of the frameworks.

Table 1. Textural properties of the prepared adsorbents and CO2 adsorption results.

12
Adsorption Amine
Surface Total Pore
Pore diameter b PEI c Nd capacity efficiency
Adsorbent area volume
(nm) (wt.%) (wt.%) (mg CO2/g (mol
(m2/g) (cm3/g)a
ads.) CO2/mol N)
SBA-15 604 1.192 7.6 0.0 0.0 5.0 N/A e
SBA-11 356 0.792 3.8 , 7.7, 12.9 0.0 0.0 0.0 N/A
SiO2 (d) 256 0.460 4.2 0.0 0.0 1.5 N/A
PEI(20)/SBA-15 248 0.698 7.1 19.7 3.3 39.5 0.38
PEI(40)/SBA-15 100 0.256 5.7 39.1 9.3 61.6 0.21
PEI(60)/SBA-15 8 0.011 3.4* 55.9 16.4 49.7 0.10

of
PEI(20)/SBA-11 110 0.534 6.0, 13.3 16.7 2.5 28.8 0.37
PEI(40)/SBA-11 16 0.116 6.8, 20.1 39.5 10.9 32.9 0.10

ro
PEI(60)/SBA-11 4 0.004 17.4 55.8 18.6 26.5 0.05
PEI(20)/ SiO2 (d) 65 0.175 3.1 , 5.4 18.8 2.8 26.4 0.30
PEI(40)/ SiO2 (d) 5 0.009 3.7* 38.2
-p 9.9 0.7 0.00
a b e
Total pore volume at P/Po = 0.99. Calculated by the BJH method from the desorption branch.
d e
re
Determined by TGA-air. N is the nitrogen content determined by elemental analysis. Not
applicable.* Not reliable information due to excessive pore blockage (see supplementary information).
lP

3.2. CO2 adsorption tests

Isotherms presented in Fig. 4 provide a comparison of the CO2 adsorption dynamics over the
RHA-derived silica adsorbents. Bare silicas exhibited negligible CO2 adsorption during the evaluated
na

exposure time (100 min), indicating that neither physical nor chemical adsorption was significant in
these kinds of materials. The adsorption efficiency of the silica supports was greatly improved by
ur

incorporating amine functionalities through impregnation with branched PEI polymer. In the amine-
modified silicas, the CO2 capture was characterized by two regions with different adsorption rates.
During the first minute of exposure, the CO2 adsorption was extremely rapid, revealing the high
Jo

affinity of the incorporated RNH2 and R2NH functionalities towards CO2. Later, the adsorption rate
was gradually decreased until the maximum adsorption capacity was reached. CO2 adsorption over
amine-based adsorbents is well known to occur through a mixed mechanism involving both chemical
and physical adsorption [54,55]. During the early exposure times, the adsorption is dominated by the
chemical interaction between the amine sites and CO2, resulting in fast adsorption rates. Once all the
adsorption sites are saturated, the adsorbed CO2 begins to fill the pores through a multilayer adsorption
process and the adsorption rate is gradually decreased [50].
13
of
ro
-p
re
Fig. 4. (A-C) CO2 adsorption isotherms of the RHA-derived adsorbents measured at 40 ºC under
lP

ambient pressure using a feed gas composition of 44.4% CO2:55.6% N2; (D-F) CO2 adsorption rates
at early exposure times.
na

Additionally, these two different adsorption rate regimes have been considered to result from
changes in the CO2 mass transfer resistance through the porous structure of the adsorbents [56]. In the
early adsorption period, there is a high concentration of amine active sites exposed to the feed gas and
ur

the mass transfer resistance is low, therefore CO2 can react rapidly. As the exposure time continues,
the active amine sites are progressively occupied by CO2 forming ammonium carbamates under dry
Jo

conditions. As a consequence, the mass diffusion of the incoming CO2 is decreased thereby reducing
the contact probability with the less accessible unreacted amine sites and resulting in slow adsorption
rates.

The porous structure of the support played an important role in the CO2 mass transfer resistance
and the adsorption rate performance. From Fig. 4 (D-F) it was found that SBA-15 based adsorbents
exhibited the highest adsorption rates compared to the PEI-impregnated SBA-11 and SiO2 (d). This
fact can be ascribed to the uniform open-ended cylindrical mesoporous structure of SBA-15, which
14
provides an easy pathway for the diffusion of CO2 molecules towards the active amine sites during
adsorption. In contrast, SBA-11 is expected to present higher CO2 mass transfer resistance due to its
interconnected three-dimensional porous system triggering slow adsorption rates [57]. This effect was
more noticeable in the case of the SiO2 (d) silica having a disordered pore network structure.

The effect of PEI loading on the CO2 adsorption capacity of the RHA-derived silica adsorbents is
shown in Fig. 5A. As expected intuitively, the increase in PEI content from 20 to 40 wt.% caused the
rise of the CO2 adsorption capacity of the SBA-11 and SBA-15 adsorbents from 28.8 to 32.9 mg CO2/g
ads and from 39.5 to 61.6 mg CO2/g ads, respectively. This improvement in the CO2 adsorption is due
to the fact that the concentration of amine functional groups also increases. Nonetheless, when the
adsorbents were loaded with 60 wt% PEI, their adsorption capacity decreased because at this critical

of
PEI content there is a complete pore filling. In the case of the disordered SiO2 (d), the adsorption
capacity was dramatically decreased at PEI loadings above 20 wt.% owing to its poor textural features.

ro
The reduction in CO2 adsorption capacity with a further increase in PEI concentration could be
understood as the result of two combined effects. On the one hand, the pore filling of the silica supports
affects the mass transport behavior of CO2 to reach the less accessible amine sites in the porous
-p
network. On the other hand, excessive PEI loadings promote the agglomeration of the polymer along
the pore channels, reducing the effective interaction of the active amine sites with the CO2 in the feed
re
gas.
lP
na
ur
Jo

Fig. 5. (A) CO2 adsorption capacity of RHA-derived silica adsorbents at different PEI loadings, and
(B) Influence of temperature on CO2 adsorption capacity of PEI(20)/SBA-15.

In this regard, since not all amine groups are able to react with CO2, a crucial parameter to consider
when evaluating the adsorbents performance is the amine efficiency. This parameter is defined as the
15
adsorbed amount of CO2 normalized by the nitrogen content as a function of the amine loaded in the
material (mol CO2/mol N) [56]. Considering only the chemical interaction between CO2 and amine
groups, the maximum efficiency of a supported amine material under dry conditions is set to 0.50 due
to the 2:1 stoichiometry (Eq. 1). The amine efficiency values of the prepared adsorbents are
summarized in Table 1. The silica supports impregnated at 20 wt.% PEI exhibited a high amine
utilization close to the maximum theoretical value (CO2/mol N = 0.38 (SBA-15), 0.37 (SBA-11), 0.30
(SiO2 (d)), indicating a high amine dispersion on the materials. In all the series of adsorbents, the amine
efficiency was gradually decreased with an increase in the PEI content due to the agglomeration of a
fraction of amine sites, which will no longer be in contact with the feed gas. The interaction between
PEI molecules and the pore walls of the silica supports has also been reported as a limiting factor for
the amines utilization [38,52]. From the FTIR analysis shown in Fig. S5 it was observed that the

of
intensity of Si-O-Si and Si-OH vibrations were decreased as PEI content increases as a result of the
hydrogen bond interaction between them [58]. As a consequence, the affinity of these amine sites for

ro
CO2 adsorption is restricted.

Considering the nature of the silica supports, the amine efficiency was increased in the order
-p
SiO2 (d) < SBA-11 < SBA-15. This trend demonstrates the importance of using pore structure-directing
templates during the synthesis of the RHA-derived silicas to modulate the pore features of the supports
re
and to enhance their performance during CO2 adsorption. The superior CO2 capacity and amine
efficiency of the SBA-15 based adsorbents are ascribed to its outstanding textural properties. Wide
lP

pores reduce the CO2 mass transfer resistance, while large pore volumes and high surface areas
improve the PEI distribution through the pores and trigger a high amine utilization.

PEI(20)/SBA-15 was selected to evaluate the influence of the adsorption temperature on the CO2
na

capacity (Fig. 5B). A slight increase in the adsorbed amount of CO2 was observed when the adsorption
temperature was raised from 40 ºC to 60 ºC. This partial improvement is attributed to the decrease in
ur

the viscosity of PEI, which distributes more uniformly in the pore paths. This behavior exposes the
amine active sites to the feed gas to a higher extent and, therefore, the amine activity is enhanced [30].
Jo

After reaching a maximum adsorption capacity of 45.4 mg CO2/g adsorbent at 60 ºC, the further
increase in temperature led to a progressive decrease in the adsorbed amount of CO2.

To further understand the effect of the PEI loading and the adsorption temperature on the CO2
capture performance of PEI/SBA-15, a kinetic study was conducted on the experimental data results.
The mechanism of the CO2 adsorption over supported amine materials is a complex phenomenon
involving both chemisorption and physisorption. Among the current available kinetic models used to
explain the CO2 adsorption with amines, the fractional-order remains one of the most suitable

16
compared with the pseudo-first order, which mainly involves physical adsorption, and the pseudo-
second order kinetic models related to chemical adsorption processes [18]. Heydari-Gorji and Sayari
[59] recently proposed a semiempirical fractional-order kinetic equation to describe the adsorption rate
of CO2 on amine-functionalized PE-MCM-41 adsorbents, which has been successfully applied in
several related works [50,54,60,61]. In this model, the adsorption rate is assumed to be proportional to
the nth power of the driving force and mth power of the adsorption time, as illustrated in Eq. 3:

1
𝑛−1 1−𝑛
𝑞𝑡 = 𝑞𝑒 − (𝑞𝑒1−𝑛 + 𝑘𝑛 𝑡 𝑚 ) (𝟑)
𝑚

Where, 𝑞𝑒 and 𝑞𝑡 stand for the adsorption capacity (mg/g) at equilibrium and at any time t,
respectively; 𝑘𝑛 is a kinetic constant involving the combination of several factors associated with

of
multiple adsorption pathways [62]; and m and n are fractional-order constants that reflect the effect of
the diffusion resistance and driving force during adsorption.

ro
The evolution of the CO2 adsorption capacity over time of the PEI-impregnated SBA-15 sorbents
and the respective fitted curves obtained by applying the kinetic model are presented in Fig. S6. Table
-p
2 summarizes the values of the kinetic parameters along with the coefficients of determination (R2)
and Mean Absolute Percentage Errors (MAPE) relative to the fitting. The mathematical fitting was
re
high, attaining elevated coefficient of determination values (R2>0.969) and low percentage errors
(MAPE<1.84%). In all cases, the calculated equilibrium capacity (𝑞𝑒 ) was quite close to the
lP

experimental value, which confirms that the fractional-order kinetic model describes well enough the
experimental data results.

Table 2. Parameter values of the fitted fractional-order kinetic model.


na

kn
qe MAPE%
Adsorbent T (ºC) (gn−1/minm m n R2 a
ur

(mg/g ads.) n−1


mg )
PEI(60)/SBA-15 40 56.0 0.732 0.30 0.50 0.994 1.16
Jo

PEI(40)/SBA-15 40 63.0 0.617 0.26 0.70 0.981 0.94


PEI(20)/SBA-15 40 39.9 0.449 1.52 1.47 0.989 0.68
PEI(20)/SBA-15 60 45.7 5.5E-06 2.47 4.43 0.969 1.84
PEI(20)/SBA-15 80 39.4 6.5E-06 2.89 6.31 0.982 0.56
PEI(20)/SBA-15 100 23.8 1.915 1.69 1.49 0.970 0.76
PEI(20)/SBA-15 120 22.1 0.015 1.06 1.45 0.999 1.58

17
1
a
MAPE (mean absolute percentage error)= ∑𝑁
𝑖=1 |(𝑞𝑡𝑒𝑥𝑝 − 𝑞𝑡𝑝𝑟𝑒𝑑 )/𝑞𝑡𝑒𝑥𝑝 |, 𝑞𝑡 is the CO2 sorption
𝑁

capacity at a given time t and N is the total number of experimental points.

The kinetic parameters m and n were affected by PEI loading and the adsorption temperature. The
value of m reflects how fast the adsorption process occurs considering diffusional limitations, while n
determines the pseudo-order of the reaction with respect to the adsorption driving force [59]. A
decrease in both m and n parameters was found with the increase in the PEI loading, confirming the
assumption that high PEI contents slow down the CO2 adsorption rate due to the decrease in the driving
force (accessible amine active sites) and the increase in the CO2 mass transfer resistance. Conversely,
low PEI loadings promoted the decrease in kn values suggesting that the reaction between CO2 and the
amine sites requires less energy and, therefore, the chemical adsorption occurs more easily under these

of
conditions [61].

On the other hand, when the adsorption temperature was increased from 40 ºC to 80 ºC, both m

ro
and n increased until reach a maximum at 80 ºC. Afterward, those values decreased with a further
increase in the temperature up to 120 ºC. This behavior indicates a change in the governance between
-p
chemical kinetics and thermodynamics of the CO2 adsorption process [63]. The large values of n at
low temperatures indicate that the adsorption rate is strongly dependent on the driving force. Under
re
low temperatures, the adsorption is essentially dominated by the dispersion of amine active sites and
the CO2 mass transfer through the porous network of the adsorbent. An increase in the temperature
lP

under this regime enhances the kinetic energy of CO2 molecules and decreases the diffusional
resistance, in consequence, the probability to reach the less accessible amine sites is favored. In Fig.
5B, the adsorbent attains its maximal CO2 adsorption capacity at 60 ºC, corresponding to the optimal
na

temperature at which the dynamics and the thermodynamics of the adsorption process reach the
equilibrium. However, when the adsorption temperature is higher than this optimal temperature, a
kinetic threshold is reached, and CO2 adsorption becomes largely controlled by thermodynamics. For
ur

this reason, the further increase in temperature led to a progressive diminishing in the amount of CO2
adsorbed, as a result of the exothermic nature of the reaction of CO2 with amines [30,55].
Jo

Finally, to verify the regeneration performance and cyclic adsorption properties of the
PEI(20)/SBA-15 adsorbent, four adsorption-desorption cycles were carried out at 40 ºC under
exposure to 44.4%CO2/55.6%N2 and ambient pressure (Fig. 6). The CO2 adsorption capacity was
gradually decreased after use, indicating that the adsorbent could not be completely regenerated. This
could be due to the partial evaporation of amine active sites or because not all the pre-adsorbed CO2
was released during the desorption process. However, after operating three isothermal cycles, only a
decrease of 6.8 % of the initial adsorption capacity was recorded. Even in the fourth cycle, 69% of the
18
starting adsorption capacity could be maintained, confirming the recyclability of the adsorbent.
Adsorbent regeneration is a crucial step in the CO2 capture process since not only determines the
stability of the sorbent when subjected to multiple adsorption cycles but also because directly
influences the recovery of the adsorbed CO2 for further utilization. Similar to the adsorption step, the
desorption of CO2 from amine-supported adsorbents involves both chemical and physical phenomena.
During isothermal regeneration under N2-stripping currents, the physically adsorbed CO2 is first
desorbed and then the molecules covalently anchored to the amine groups are released [60]. The latter
implies the thermal decomposition of the ammonium-carbamate pairs previously formed in the
adsorption step under dry conditions following the reverse reaction of that described in Eq. 1 [64].

of
ro
-p
re
lP

Fig. 6. Stability of PEI(20)/SBA-15 under repeated adsorption cycles at 40 ºC.


na

4. CONCLUSIONS

The CO2 adsorption behavior of a series of amine-functionalized silicas with a 2D hexagonal, 3D


ur

cubic or disordered pore system was studied in this work. The pore structure of the silicas and the
incorporated amount of amine strongly influences the performance of the adsorbents. Ordered straight
Jo

mesopores of the two-dimensional system provide higher adsorption rate, CO2 adsorption capacity and
amine efficiency compared to the interconnected three-dimensional and disordered porous structures,
which hinder the CO2 mass transfer and the accessibility to unreacted amine adsorption sites.

Efficient pore filling is a crucial factor in order to control the amine dispersion and also to avoid
the excessive pore blockage of the adsorbents. High PEI loading supplies more amine sites able to
react but also increases CO2 diffusional barriers due to plug-like agglomeration throughout the pores.

19
Therefore, a balance between the porosity features of the support and the extent of amine loading must
be considered to develop novel CO2 adsorbents.

The CO2 adsorption process is dominated by kinetic and thermodynamic regimes to different
extents depending on the operating temperature. At low temperatures, the adsorption is mainly limited
by the accessibility of amine active sites and the CO2 mass transfer through the porous network.
However, once a kinetic threshold is reached at a certain temperature, the CO2 adsorption becomes
largely controlled by thermodynamics being less favorable based on its exothermic nature.

Conflict of Interest

We know of no conflicts of interest associated with this publication, and there has been no significant financial
support for this work that could have influenced its outcome. As Corresponding Author, I confirm that the

of
manuscript has been read and approved for submission by all the named authors.

ro
Declaration of interests

-p
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.
re
lP

ACKNOWLEDGMENTS
na

The authors acknowledge the financial support from the program “Jóvenes Investigadores e
Innovadores No. 812” of the Colombian Ministry of Science, Technology and Innovation. Dr. Juan D.
ur

Martínez (UPB-Medellín) is acknowledged for supplying the RHA. Laboratorio de Ciencias Térmicas
at Instituto Tecnológico Metropolitano (ITM) and Universidad de Antioquia (UdeA) are
acknowledged for providing the equipment to perform the experiments and characterization.
Jo

REFERENCES

[1] A. Mohmmed, Z. Li, A. Olushola Arowolo, H. Su, X. Deng, O. Najmuddin, Y. Zhang, Driving
factors of CO2 emissions and nexus with economic growth, development and human health in
the Top Ten emitting countries, Resour. Conserv. Recycl. 148 (2019) 157–169.
https://doi.org/10.1016/j.resconrec.2019.03.048.
[2] A. Valadkhani, R. Smyth, J. Nguyen, Effects of primary energy consumption on CO2 emissions
20
under optimal thresholds: Evidence from sixty countries over the last half century, Energy Econ.
80 (2019) 680–690. https://doi.org/10.1016/j.eneco.2019.02.010.
[3] J. Yan, Z. Zhang, Carbon Capture, Utilization and Storage (CCUS), Appl. Energy. 235 (2019)
1289–1299. https://doi.org/10.1016/j.apenergy.2018.11.019.
[4] M. Romero-Sáez, L.Y. Jaramillo, W. Henao, U. de la Torre, Nanomaterials for CO2
Hydrogenation, in: S. Rajendran, M. Naushad, K. Raju, R. Boukherroub (Eds.), Emerg.
Nanostructured Mater. Energy Environ. Sci., 2019: pp. 173–214. https://doi.org/10.1007/978-
3-030-04474-9_4.
[5] F. Marocco Stuardi, F. MacPherson, J. Leclaire, Integrated CO2 capture and utilization: A
priority research direction, Curr. Opin. Green Sustain. Chem. 16 (2019) 71–76.
https://doi.org/10.1016/j.cogsc.2019.02.003.

of
[6] T. Lockwood, A Compararitive Review of Next-generation Carbon Capture Technologies for
Coal-fired Power Plant, Energy Procedia. 114 (2017) 2658–2670.

ro
https://doi.org/10.1016/j.egypro.2017.03.1850.
[7] A. Modak, S. Jana, Advancement in porous adsorbents for post-combustion CO2 capture,
Microporous Mesoporous Mater.
https://doi.org/10.1016/j.micromeso.2018.09.018.
-p 276 (2019) 107–132.
re
[8] D.W.F. Brilman, R. Veneman, Capturing atmospheric CO2 using supported amine sorbents,
Energy Procedia. 37 (2013) 6070–6078. https://doi.org/10.1016/j.egypro.2013.06.536.
lP

[9] B. Dutcher, M. Fan, A.G. Russell, Amine-based CO2 capture technology development from the
beginning of 2013-A review, ACS Appl. Mater. Interfaces. 7 (2015) 2137–2148.
https://doi.org/10.1021/am507465f.
na

[10] C. Chen, S. Zhang, K.H. Row, W.S. Ahn, Amine–silica composites for CO2 capture: A short
review, J. Energy Chem. 26 (2017) 868–880. https://doi.org/10.1016/j.jechem.2017.07.001.
[11] R. Sanz, G. Calleja, A. Arencibia, E.S. Sanz-Pérez, CO2 uptake and adsorption kinetics of pore-
ur

expanded SBA-15 double-functionalized with amino groups, Energy and Fuels. 27 (2013)
7637–7644. https://doi.org/10.1021/ef4015229.
Jo

[12] D.V. Quang, A. Dindi, M.R.M. Abu-Zahra, One-Step Process Using CO2 for the Preparation of
Amino-Functionalized Mesoporous Silica for CO2 Capture Application, ACS Sustain. Chem.
Eng. 5 (2017) 3170–3178. https://doi.org/10.1021/acssuschemeng.6b02961.
[13] K. Li, J. Jiang, S. Tian, X. Chen, Polyethyleneimine–nano silica composites: a low- cost and
promising adsorbent for CO2 capture, J. Mater. Chem. A. 3 (2015) 2166–2175.
https://doi.org/10.1039/C4TA04275A.
[14] R. Chatti, A.K. Bansiwal, J.A. Thote, V. Kumar, P. Jadhav, S.K. Lokhande, R.B. Biniwale,

21
N.K. Labhsetwar, S.S. Rayalu, Amine loaded zeolites for carbon dioxide capture: Amine
loading and adsorption studies, Microporous Mesoporous Mater. 121 (2009) 84–89.
https://doi.org/10.1016/j.micromeso.2009.01.007.
[15] S. Zhang, C. Chen, W.S. Ahn, Recent progress on CO2 capture using amine-functionalized
silica, Curr. Opin. Green Sustain. Chem. 16 (2019) 26–32.
https://doi.org/10.1016/j.cogsc.2018.11.011.
[16] T. Ghanbari, F. Abnisa, W.M.A. Wan Daud, A review on production of metal organic
frameworks (MOF) for CO2 adsorption, Sci. Total Environ. 707 (2020) 135090.
https://doi.org/10.1016/j.scitotenv.2019.135090.
[17] M.S. Raja Shahrom, A.R. Nordin, C.D. Wilfred, The improvement of activated carbon as CO2
adsorbent with supported amine functionalized ionic liquids, J. Environ. Chem. Eng. 7 (2019)

of
103319. https://doi.org/10.1016/j.jece.2019.103319.
[18] A.A. Azmi, M.A.A. Aziz, Mesoporous adsorbent for CO2 capture application under mild

ro
condition: A review, J. Environ. Chem. Eng. 7 (2019) 103022.
https://doi.org/10.1016/j.jece.2019.103022.
-p
[19] U. Zulfiqar, T. Subhani, S. Wilayat Husain, Synthesis of silica nanoparticles from sodium
silicate under alkaline conditions, J. Sol-Gel Sci. Technol. 77 (2016) 753–758.
re
https://doi.org/10.1007/s10971-015-3950-7.
[20] S. Affandi, H. Setyawan, S. Winardi, A. Purwanto, R. Balgis, A facile method for production
lP

of high-purity silica xerogels from bagasse ash, Adv. Powder Technol. 20 (2009) 468–472.
https://doi.org/10.1016/j.apt.2009.03.008.
[21] United States Department of Agriculture-USDA, World agricultural production, Foreign Agric.
na

Serv. Glob. Mark. Anal. (2020) 1–40.


https://apps.fas.usda.gov/psdonline/circulars/production.pdf.
[22] V.P. Della, I. Kühn, D. Hotza, Rice husk ash as an alternate source for active silica production,
ur

Mater. Lett. 57 (2002) 818–821. https://doi.org/10.1016/S0167-577X(02)00879-0.


[23] E.R. Abaide, M. V. Tres, G.L. Zabot, M.A. Mazutti, Reasons for processing of rice coproducts:
Jo

Reality and expectations, Biomass and Bioenergy. 120 (2019) 240–256.


https://doi.org/10.1016/j.biombioe.2018.11.032.
[24] N. Soltani, A. Bahrami, M.I. Pech-Canul, L.A. González, Review on the physicochemical
treatments of rice husk for production of advanced materials, Chem. Eng. J. 264 (2015) 899–
935. https://doi.org/10.1016/j.cej.2014.11.056.
[25] C. Gérardin, J. Reboul, M. Bonne, B. Lebeau, Ecodesign of ordered mesoporous silica
materials, Chem. Soc. Rev. 42 (2013) 4217–4255. https://doi.org/10.1039/c3cs35451b.

22
[26] Y. Han, D. Zhang, Ordered mesoporous silica materials with complicated structures, Curr. Opin.
Chem. Eng. 1 (2012) 129–137. https://doi.org/10.1016/j.coche.2011.11.001.
[27] D. Li, X. Guan, J. Song, Y. Di, D. Zhang, X. Ge, L. Zhao, F.S. Xiao, Highly efficient synthesis
of ordered mesoporous silica materials with controllable microporosity using surfactant
mixtures as templates, Colloids Surfaces A Physicochem. Eng. Asp. 272 (2006) 194–202.
https://doi.org/10.1016/j.colsurfa.2005.07.027.
[28] E.S. Sanz-Pérez, A. Arencibia, G. Calleja, R. Sanz, Tuning the textural properties of HMS
mesoporous silica. Functionalization towards CO2 adsorption, Microporous Mesoporous Mater.
260 (2018) 235–244. https://doi.org/10.1016/j.micromeso.2017.10.038.
[29] R. Sanz, G. Calleja, A. Arencibia, E.S. Sanz-Pérez, CO2 capture with pore-expanded MCM-41
silica modified with amino groups by double functionalization, Microporous Mesoporous

of
Mater. 209 (2015) 165–171. https://doi.org/10.1016/j.micromeso.2014.10.045.
[30] R. Kishor, A.K. Ghoshal, Amine-modified mesoporous silica for CO2 adsorption: The role of

ro
structural parameters, Ind. Eng. Chem. Res. 56 (2017) 6078–6087.
https://doi.org/10.1021/acs.iecr.7b00890.
-p
[31] D. Wang, X. Wang, X. Ma, E. Fillerup, C. Song, Three-dimensional molecular basket sorbents
for CO2 capture: Effects of pore structure of supports and loading level of polyethylenimine,
re
Catal. Today. 233 (2014) 100–107. https://doi.org/10.1016/j.cattod.2014.01.038.
[32] C.H. Lee, D.H. Hyeon, H. Jung, W. Chung, D.H. Jo, D.K. Shin, S.H. Kim, Effects of pore
lP

structure and PEI impregnation on carbon dioxide adsorption by ZSM-5 zeolites, J. Ind. Eng.
Chem. 23 (2015) 251–256. https://doi.org/10.1016/j.jiec.2014.08.025.
[33] Y. Le, D. Guo, B. Cheng, J. Yu, Amine-functionalized monodispersed porous silica
na

microspheres with enhanced CO2 adsorption performance and good cyclic stability, J. Colloid
Interface Sci. 408 (2013) 173–180. https://doi.org/10.1016/j.jcis.2013.07.014.
[34] P. Zhao, G. Zhang, Y. Sun, Y. Xu, CO2 adsorption behavior and kinetics on amine-
ur

functionalized composites silica with trimodal nanoporous structure, Energy Fuels 31 (2017)
12508–12520. https://doi.org/10.1021/acs.energyfuels.7b02292.
Jo

[35] J. Yu, Y. Le, B. Cheng, Fabrication and CO2 adsorption performance of bimodal porous silica
hollow spheres with amine-modified surfaces, RSC Adv. 2 (2012) 6784–6791.
https://doi.org/10.1039/c2ra21017g.
[36] Y.G. Ko, H.J. Lee, J.Y. Kim, U.S. Choi, Hierarchically porous aminosilica monolith as a CO2
adsorbent, ACS Appl. Mater. Interfaces. 6 (2014) 12988–12996.
https://doi.org/10.1021/am5029022.
[37] K. Li, J. Jiang, F. Yan, S. Tian, X. Chen, The influence of polyethyleneimine type and molecular

23
weight on the CO2 capture performance of PEI-nano silica adsorbents, Appl. Energy. 136 (2014)
750–755. https://doi.org/10.1016/j.apenergy.2014.09.057.
[38] A. Heydari-Gorji, Y. Belmabkhout, A. Sayari, Polyethylenimine-impregnated mesoporous
silica: Effect of amine loading and surface alkyl chains on CO2 adsorption, Langmuir. 27 (2011)
12411–12416. https://doi.org/10.1021/la202972t.
[39] Z.L. Liu, Y. Teng, K. Zhang, Y. Cao, W.P. Pan, CO2 adsorption properties and thermal stability
of different amine-impregnated MCM-41 materials, J. Fuel Chem. Technol. 41 (2013) 469–476.
https://doi.org/10.1016/S1872-5813(13)60025-0.
[40] J.D. Martínez, T. Pineda, J.P. López, M. Betancur, Assessment of the rice husk lean-combustion
in a bubbling fluidized bed for the production of amorphous silica-rich ash, Energy. 36 (2011)
3846–3854. https://doi.org/10.1016/j.energy.2010.07.031.

of
[41] L.Y. Jaramillo, K. Arango-Benítez, W. Henao, E. Vargas, G. Recio-Sánchez, M. Romero-Sáez,
Synthesis of ordered mesoporous silicas from rice husk with tunable textural properties, Mater.

ro
Lett. 257 (2019). https://doi.org/10.1016/j.matlet.2019.126749.
[42] T.H. Liou, C.C. Yang, Synthesis and surface characteristics of nanosilica produced from alkali-
-p
extracted rice husk ash, Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 176 (2011) 521–
529. https://doi.org/10.1016/j.mseb.2011.01.007.
re
[43] G. Calleja, R. Sanz, A. Arencibia, E.S. Sanz-Pérez, Influence of drying conditions on amine-
functionalized SBA-15 as adsorbent of CO2, Top. Catal. 54 (2011) 135–145.
lP

https://doi.org/10.1007/s11244-011-9652-7.
[44] L. Chen, J. Xu, W.H. Zhang, J.D. Holmes, M.A. Morris, Syntheses of complex mesoporous
silicas using mixtures of nonionic block copolymer surfactants: Understanding formation of
na

different structures using solubility parameters, J. Colloid Interface Sci. 353 (2011) 169–180.
https://doi.org/10.1016/j.jcis.2010.09.043.
[45] J. Wang, H. Ge, W. Bao, Synthesis and characteristics of SBA-15 with thick pore wall and high
ur

hydrothermal stability, Mater. Lett. 145 (2015) 312–315.


https://doi.org/10.1016/j.matlet.2015.01.113.
Jo

[46] D. Zhao, Q. Huo, J. Feng, B.F. Chmelka, G.D. Stucky, Nonionic triblock and star diblock
copolymer and oligomeric sufactant syntheses of highly ordered, hydrothermally stable,
mesoporous silica structures, J. Am. Chem. Soc. 120 (1998) 6024–6036.
https://doi.org/10.1021/ja974025i.
[47] R. Guillet-Nicolas, R. Ahmad, K.A. Cychosz, F. Kleitz, M. Thommes, Insights into the pore
structure of KIT-6 and SBA-15 ordered mesoporous silica-recent advances by combining
physical adsorption with mercury porosimetry, New J. Chem. 40 (2016) 4351–4360.

24
https://doi.org/10.1039/c5nj03466c.
[48] X.Y. Yang, L.H. Chen, Y. Li, J.C. Rooke, C. Sanchez, B.L. Su, Hierarchically porous materials:
Synthesis strategies and structure design, Chem. Soc. Rev. 46 (2017) 481–558.
https://doi.org/10.1039/c6cs00829a.
[49] A. Boonpoke, S. Chiarakorn, N. Laosiripojana, A. Chidthaisong, Enhancement of carbon
dioxide capture by amine-modified Rice Husk mesoporous material, Environ. Prog. Sustain.
Energy. 35 (2016) 1716–1723. https://doi.org/10.1002/ep.12423.
[50] Y. Wang, T. Du, Z. Qiu, Y. Song, S. Che, X. Fang, CO2 adsorption on polyethylenimine-
modified ZSM-5 zeolite synthesized from rice husk ash, Mater. Chem. Phys. 207 (2018) 105–
113. https://doi.org/10.1016/j.matchemphys.2017.12.040.
[51] I. Yudovin-Farber, N. Beyth, E.I. Weiss, A.J. Domb, Antibacterial effect of composite resins

of
containing quaternary ammonium polyethyleneimine nanoparticles, J. Nanoparticle Res. 12
(2010) 591–603. https://doi.org/10.1007/s11051-009-9628-8.

ro
[52] A. Holewinski, M.A. Sakwa-Novak, C.W. Jones, Linking CO2 sorption performance to polymer
morphology in aminopolymer/silica composites through neutron scattering, J. Am. Chem. Soc.
-p
137 (2015) 11749–11759. https://doi.org/10.1021/jacs.5b06823.
[53] X. Guo, L. Ding, K. Kanamori, K. Nakanishi, H. Yang, Functionalization of hierarchically
re
porous silica monoliths with polyethyleneimine (PEI) for CO2 adsorption, Microporous
Mesoporous Mater. 245 (2017) 51–57. https://doi.org/10.1016/j.micromeso.2017.02.076.
lP

[54] Y. Wang, T. Du, Y. Song, S. Che, X. Fang, L. Zhou, Amine-functionalized mesoporous ZSM-
5 zeolite adsorbents for carbon dioxide capture, Solid State Sci. 73 (2017) 27–35.
https://doi.org/10.1016/j.solidstatesciences.2017.09.004.
na

[55] G. Zhang, P. Zhao, L. Hao, Y. Xu, Amine-modified SBA-15(P): A promising adsorbent for CO2
capture, J. CO2 Util. 24 (2018) 22–33. https://doi.org/10.1016/j.jcou.2017.12.006.
[56] M.A. Sakwa-Novak, S. Tan, C.W. Jones, Role of Additives in Composite PEI/Oxide CO 2
ur

Adsorbents: Enhancement in the amine efficiency of supported PEI by PEG in CO2 capture
from simulated ambient air, ACS Appl. Mater. Interfaces. 7 (2015) 24748–24759.
Jo

https://doi.org/10.1021/acsami.5b07545.
[57] T.J. Rottreau, C.M.A. Parlett, A.F. Lee, R. Evans, Diffusion NMR characterization of catalytic
silica supports: A tortuous path, J. Phys. Chem. C. 121 (2017) 16250–16256.
https://doi.org/10.1021/acs.jpcc.7b02929.
[58] Z. Ghazali, N. Suhaili, M.N.A. Tahari, M.A. Yarmo, N.H. Hassan, R. Othaman, Impregnating
deep eutectic solvent choline chloride: Urea: Polyethyleneimine onto mesoporous silica gel for
carbon dioxide capture, J. Mater. Res. Technol. (2020).

25
https://doi.org/10.1016/j.jmrt.2020.01.073.
[59] A. Heydari-Gorji, A. Sayari, CO2 capture on polyethylenimine-impregnated hydrophobic
mesoporous silica: Experimental and kinetic modeling, Chem. Eng. J. 173 (2011) 72–79.
https://doi.org/10.1016/j.cej.2011.07.038.
[60] Q. Liu, J. Shi, S. Zheng, M. Tao, Y. He, Y. Shi, Kinetics studies of CO 2 adsorption/desorption
on amine-functionalized multiwalled carbon nanotubes, Ind. Eng. Chem. Res. 53 (2014) 11677–
11683. https://doi.org/10.1021/ie502009n.
[61] J. Ouyang, C. Zheng, W. Gu, Y. Zhang, H. Yang, S.L. Suib, Textural properties determined
CO2 capture of tetraethylenepentamine loaded SiO2 nanowires from Α-sepiolite, Chem. Eng. J.
337 (2018) 342–350. https://doi.org/10.1016/j.cej.2017.12.109.
[62] R. Serna-Guerrero, A. Sayari, Modeling adsorption of CO2 on amine-functionalized

of
mesoporous silica. 2: Kinetics and breakthrough curves, Chem. Eng. J. 161 (2010) 182–190.
https://doi.org/10.1016/j.cej.2010.04.042.

ro
[63] A. Zhao, A. Samanta, P. Sarkar, R. Gupta, Carbon dioxide adsorption on amine-impregnated
mesoporous SBA-15 sorbents: Experimental and kinetics study, Ind. Eng. Chem. Res. 52 (2013)
6480–6491. https://doi.org/10.1021/ie3030533. -p
[64] Z. Sun, M. Fan, M. Argyle, Desorption kinetics of the monoethanolamine/macroporous TiO2-
re
based CO2 separation process, Energy and Fuels. 25 (2011) 2988–2996.
https://doi.org/10.1021/ef200556j.
lP
na
ur
Jo

26

You might also like