Sie sind auf Seite 1von 9

Fuel 206 (2017) 10–18

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Modified carbon nanotubes/tetraethylenepentamine for CO2 capture


Maryam Irani, Andrew T. Jacobson, Khaled A.M. Gasem, Maohong Fan ⇑
Department of Chemical and Petroleum Engineering, University of Wyoming, Laramie, WY 82071, USA

h i g h l i g h t s

 Modified carbon nanotubes (MCNTs) were prepared using a KOH reagent.


 MCNTs/tetraethylenepentamine was prepared as a CO2 sorbent.
 Under optimal conditions, the CO2 sorption capacity reached 5 mmol CO2/g.
 High CO2 desorption rate of the sorbent can considerably reduce CO2 capture cost.

a r t i c l e i n f o a b s t r a c t

Article history: In this work, a CO2 sorbent was prepared by immobilizing tetraethylenepentamine (TEPA) onto modified
Received 1 April 2017 carbon nanotubes. Modification of carbon nanotubes (CNTs) using a KOH reagent was done to increase
Received in revised form 28 May 2017 the surface area and pore volume of the CNTs. The prepared sorbents were characterized using X-ray
Accepted 29 May 2017
diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), scanning electron microscopy (SEM),
transmission electron microscopy (TEM) thermogravimetric analysis (TGA), and Brunauer-Emmett-
Teller (BET) analyses. At the optimal TEPA loading of 75 wt% on modified CNTs (MCNTs), the CO2 sorption
Keywords:
capacity reached 5 mmol CO2/g-sorbent for 10 vol% CO2 in N2 along with 1 vol% H2O at 60 °C. Kinetic and
CO2 capture
Carbon nanotubes
thermodynamic adsorption studies found activation energies for CO2 adsorption and desorption of
Sorption MCNTs/TEPA being16.2 kJ/mol and 39.9 kJ/mol, respectively. The low activation energy for CO2 desorp-
Kinetics tion using MCNTs/TEPA corresponds with a high CO2 desorption rate, resulting in a low CO2 capture cost.
Therefore, the MCNTs/TEPA sorbent has potential for application to CO2 capture from gas mixtures.
Ó 2017 Elsevier Ltd. All rights reserved.

1. Introduction According to predictions by the Intergovernmental Panel on


Climate Change (IPCC) the atmosphere may have a concentration
The issue of increasing CO2 emissions in the environment has of up to 570 ppm CO2 by the year 2100, which may result in a
received much attention because of its association with climate mean global temperature increase of around 1.9 °C and a rise in
change. Specifically, CO2 emissions contribute to the greenhouse mean sea level of 3.8 m [1].
effect; a phenomenon where greenhouse gases such as CO2, Around 85% of the world’s total energy demand is provided by
methane, nitrous oxide, water vapor, and ozone absorb and release fossil fuel thermal power plants. Fossil fuel power plants con-
thermal infrared radiation resulting in an increase of Earth’s tem- tribute approximately 40% of total CO2 emissions around the
perature [1]. The increase in concentration of greenhouse gases world. The capture of CO2 from power plant flue gas gives rise to
due to human activities has caused an increase of 0.8 °C in the around 75% of the total cost of carbon capture and storage [1].
Earth’s surface temperature within the past 100 years [2]. Excess CO2 emissions from thermal power plant flue gas can be
amounts of greenhouse gases in the atmosphere are also responsi- decreased through pre-combustion carbon capture, post-
ble for other environmental problems such as a rising water-level combustion carbon capture, and oxyfuel combustion [1,3]. The
in the world’s oceans, an increase in the number of ocean storms, focus of this work falls into the post-combustion capture category,
and an upsurge in floods around the world [1]. Among the green- which describes the process of CO2 removal from flue gas gener-
house gases, CO2 is the major source attributed to climate change. ated by thermal power plant combustion chambers that utilize
CO2 alone contributes to roughly 64% of the greenhouse effect [1]. air for combustion. These power plants produce a flue gas at atmo-
spheric pressure with an average CO2 concentration of less than
15% [1].
⇑ Corresponding author.
E-mail address: mfan@uwyo.edu (M. Fan).

http://dx.doi.org/10.1016/j.fuel.2017.05.087
0016-2361/Ó 2017 Elsevier Ltd. All rights reserved.
M. Irani et al. / Fuel 206 (2017) 10–18 11

Aqueous amine technology has been applied to CO2 absorption chased from Sigma-Aldrich. Anhydrous ethanol was purchased
in industry, but it suffers from drawbacks including high energy from Decon Labs, Inc. Tetraethylenepentamine (TEPA, technical
requirements and corrosion problems [4]. Adsorption on solid grade) and potassium hydroxide (KOH, 85%) were purchased from
adsorbents is an attractive alternative for efficient CO2 capture from Sigma-Aldrich. Hydrochloric acid (HCl, 36%–38%, wt.) was obtained
flue gases. Specifically, carbon based adsorbents are promising due from Fisher scientific, Inc. All gases used in this work were supplied
to their chemical inertness and high surface area [5]. Using solid by United States Welding, Inc.
adsorbents for CO2 capture have potential advantages compared Chemical activation of the CNTs was performed by mixing KOH
to amine scrubbing (aqueous amine technology) including a reduc- and CNTs at an activating agent/carbon ratio of 6:1 in 50 ml of
tion in regeneration energy and ease of handling. Several authors ethanol, and then stirred for 6 h at room temperature. Next the
have reviewed and compared aqueous amine technology with mixture was dried in an oven at 80 °C for at least 12 h, followed
other technologies (e.g. solid adsorbents) comprehensively [6,7]. by heat treatment at a temperature of 600 °C in N2 gas for 3 h.
Recently, studies show that modifications of porous materials The materials were then stirred in dilute HCl overnight, filtered,
by introducing basic sites can enhance CO2 adsorption capacity washed with water until a pH value of 7 was reached, and dried
and selectivity due to a strong interaction between basic groups in the oven at 100 °C. These modified CNTs are given the term
and acidic CO2. The common functional groups for modification MCNT for the remainder of this work.
of porous materials are alkaline carbonates and a number of amine The MCNTs/TEPA were prepared by wet impregnation. A speci-
groups [8]. These solid adsorbents can be prepared by either fied amount of TEPA was dissolved in 20 ml of ethanol under stir-
chemically bonding amine groups to a support, referred to as ring. Next, the MCNTs were added into the ethanol/TEPA solution
grafting [9], or immobilizing liquid amines within the pores of a and stirred for 24 h at room temperature, and then was dried in
support, known as impregnation [10]. Compared to the grafting an oven at 50 °C. The amount of TEPA and MCNTs added to the solu-
method, impregnation has the advantage of simple preparation tion was selected for preparation of MCNTs/TEPA samples with
procedures [11]. The wet impregnation of amine-based materials TEPA loadings of 50, 60, 70, 75, and 80 wt%. The unmodified
is generally a physical mixture of amine-based materials and sup- CNTs/TEPA sorbent was prepared similarly with a maximum TEPA
ports which results in a non-covalent introduction of amine-based loading of 50 wt%. TEPA loading over 50 wt% was not achievable
materials into the pores or onto the surface of the support [10]. due to the unmodified CNTs sorbent agglomerating from amine
Carbon nanotubes (CNTs) have a variety of applications due to overloading. This most likely can be attributed to the lower surface
their possession of unique properties such as topological hollow area of unmodified CNTs (CNTs without treatment with KOH) com-
tubular structures, thermal and chemical stability, and large pores pared to modified CNTs (MCNTs).
[10,12]. Su et al. [13] modified CNTs using 3-aminopropyl-
triethoxysilane (APTS), and studied CO2 adsorption capacity. The
optimal CO2 adsorption capacity was found to be 1 mmol/g for
2.2. Characterization of sorbent
15 v.% CO2 at 20 °C. Carbon nanotubes (CNTs) impregnated with
tetraethylenepentamine (TEPA) were prepared by Ye et al. [10] as
To characterize the functionalities on the sorbents, a Nicolet/
a solid amine adsorbent for CO2 capture. The maximum adsorption
iS50 spectroscope Fourier transform infrared spectrometer (FTIR)
capacity was approximately 3.87 mmol g1 for 2 v.% CO2 with 2 v.
was used with a frequency range of 4000–500 cm1. X-ray diffrac-
% H2O at 40 °C.
tion (XRD) analyses were completed with a Rigaku Smartlab X-ray
Chemical activation is a very efficient method to produce car-
diffraction system using a Cu Ka1 line (1.5406 Å) operating at
bons with high surface area. Alkaline hydroxides, such as KOH
40 kV/40 mA, with 2h ranging from 10° to 90°. A FEI Quanta 450
and NaOH, have been used as common activating agents for
field emission scanning electron microscope with a Schottky field
the preparation of activated carbons originating from a variety
emission gun at an accelerating voltage of 20 kV was applied to
of carbonaceous precursors such as coals, carbon nanotubes,
obtain scanning electron microscopy (SEM) images. A Quan-
and fibers. Generally, chemical activation using an alkaline
tachrome Autosorb IQ automated gas sorption analyzer was used
involves solid–solid or solid–liquid reactions, including the
to obtain BET (Brunauer-Emmett-Teller) surface area and BJH
hydroxide reduction and carbon oxidation, to form a porous
(Barrett-Joyner-Halenda) pore volume. A TA Instruments SDT Q
structure [14].
600 was applied for thermogravimetric analysis (TGA) under N2
In this work, KOH was chosen as an activation agent to modify
gas with a heating rate of 10 °C/min. A Tecnai G2 F20 S-Twin
multi-walled carbon nanotubes to increase the surface area and
200 kV High Resolution Transmission Electron Microscope
porosity. Several authors [10,15,16] have reported functionalized
(HRTEM) was utilized for obtaining TEM images.
unmodified carbon nanotubes (CNTs) with tetraethylenepen-
tamine (TEPA) for CO2 capture, all of which show lower CO2 cap-
ture capacity when compared to TEPA supported on KOH
modified carbon nanotubes (MCNTs/TEPA) used in this work. 2.3. CO2 sorption and regeneration experiments
The MCNTs/TEPA CO2 capture results are also compared with
experimental results from this study using a sorbent prepared Reactor preparation consisted of 50 mg of the sorbent being
from unmodified CNTs. To the best of our knowledge the use of mixed with 1000 mg of sand and loaded into a quartz tube reactor,
KOH modified carbon nanotubes as a support for the preparation followed by placement into a tube furnace. For sorption experi-
of an amine-based sorbent through impregnation has not been ments10 vol% CO2 in N2 at a flow rate of 300 mL/min was fed to
reported. the reactor. 1 vol% H2O was introduced into this inlet gas stream
by a syringe pump, and heating tape was used to vaporize the
water before entering the reactor. Each sorption test was run until
2. Experimental the outlet CO2 concentration and the inlet CO2 concentration were
equivalent. CO2 desorption was completed using the same experi-
2.1. Preparation of the MCNT/TEPA sorbents mental setup. The carrier gas, N2 gas (99.9%), at a flow rate of
300 mL/min was used for desorption of CO2 at 90 °C. A schematic
Multi walled carbon nanotubes (CNTs) with dimensions (O.D.  of the experimental set up used for CO2 separation is shown and
I.D.  L) of 10 nm ± 1 nm  4.5 nm ± 0.5 nm  3–6 nm were pur- labelled (No. 1–14) in Fig. 1.
12 M. Irani et al. / Fuel 206 (2017) 10–18

Fig. 1. Schematic diagram of CO2 separation setup: 1, CO2 cylinder; 2, N2 cylinder; 3(a) CO2 mass flow controller; 3(b), N2 mass flow controller; 4, syringe pump; 5, heat tape
temperature controller; 6, heat tape; 7, tube furnace temperature controller; 8, tube furnace; 9, quartz tube reactor; 10, quartz wool and notch block; 11, sorbent; 12, water
removal unit; 13, gas analyzer; 14, data recording system.

3. Results and discussions

3.1. Characterization

Fig. 2 shows FTIR patterns of the CNTs, MCNTs, and MCNTs/


TEPA. From the spectrum of MCNTs no changes were observed
after modification of the CNTs with KOH, suggesting no functional
groups are present. Compared to the unloaded CNTs and MCNTs,
the FTIR spectrum of MCNTs/TEPA shows additional peaks indicat-
ing the presence of TEPA. The –NH2 and –NH bending vibration of
TEPA are seen in the bands at 1520 and 1420 cm1 , respectively
[15]. Furthermore, the bands at 2870 and 2810 cm1 correspond to
C–H asymmetric and symmetric vibration peaks, respectively.
The XRD patterns of the CNTs, MCNTs, and MCNTs/TEPA are
given in Fig. 3. The strong diffraction peak at 25° and weak diffrac-
tion peaks at 42° and 77° correspond to the graphite (0 0 2), (1 0 0),
and (1 1 0) lattice planes of the CNTs, respectively [15,17]. After
modification of CNTs with KOH at 600 °C, the peak locations do Fig. 3. XRD results of the CNTs, MCNTs and MCNTs/TEPA.
not show any obvious changes indicating that the structure of
the CNTs is retained. It has been reported that KOH is transformed
into K2CO3 at 600 °C [14]. It is proposed that the concentration of MCNT (25°) overlapping the tail of the TEPA peak contributes to
K2CO3 is too low to be detected by XRD, and in part confirmed the increased broadness and intensity of the MCNTs/TEPA peak.
by EDX (image not shown in this work) indicating low concentra- No other obvious changes were seen in the peak locations of
tions of potassium. This low concentration may be due to removal MCNTs after TEPA incorporation.
of K2CO3 through the washing of MCNTs with a hydrochloric acid TGA weight loss curves of the CNTs, MCNTs, and MCNTs/TEPA are
solution. In the XRD pattern of MCNTs/TEPA, the broad diffraction shown in Fig. 4. MCNTs/TEPA displayed a mass loss of approximately
peak spanning 10° –30° with a maximum at 19.6° indicates the 7 wt% from 25 °C to 109 °C, corresponding to the evaporation of
presence of amorphous TEPA. Also, the sharp diffraction peak of

Fig. 2. FTIR spectra of CNTs, MCNTs and MCNTs/TEPA. Fig. 4. TGA curves of CNTs, MCNTs and MCNTs/TEPA with TEPA loading of 75 wt%.
M. Irani et al. / Fuel 206 (2017) 10–18 13

adsorbed water. A sharp weight loss (60%) was then observed 3.3. Effect of moisture addition
between 109 and 246 °C due to the decomposition of TEPA. This
was followed by a third step of weight loss (8%) between 246 Studying the effect of moisture on CO2 sorption capacity is cru-
and 420 °C that may be due to the loss of the residual TEPA trapped cial as CO2-containing gases, such as flue gas, typically contain
in the support. Therefore, MCNTs/TEPA are thermally stable below water vapor. The addition of 1 vol% moisture increased CO2 sorp-
110 °C which is applicable for CO2 sorption (60 °C) and desorption tion capacity by 56% (3.2–5 mmol/g) compared to dry gas adsorp-
(90 °C) in this work. tion, as seen in Fig. 9. Conversely, CO2 adsorption capacity reduced
As shown in Table 1, the surface area of the CNTs increases from with a moisture concentration greater than 1%.
90.3 to 169.8 m2/g after activation with KOH. The generation of lar- This reduction in adsorption capacity may be caused by the
ger pores after KOH activation results in an increase in surface area water molecules occupying available adsorption sites, competing
of the MCNT. After TEPA loading into the MCNTs, the surface area with CO2 [20].
and pore volumes decrease, which is an indication that TEPA has The CO2/amine reactions differ between wet and dry conditions
been successfully loaded into the MCNTs structure. Fig. 5 shows as seen below. Dry conditions facilitate the formation of carba-
TEM micrographs of unmodified CNTs and modified CNTs. It is mates between two amine groups, while wet conditions facilitate
obvious that the nanotubular morphology of CNTs is preserved the formation of bicarbonates between one CO2 molecule and
after the reaction with KOH (Fig. 5b). The high magnification image one amine group [21,22]. Two types of these amine groups exist
of MCNTs shows that the KOH activation process creates defects in within TEPA including primary (R1–NH2) and secondary (R1–
the nanotubes walls. It has been reported that these defects result NH–R2) amines. The maximum adsorption capacity was calculated
in the increase of pore volumes and surface area of the modified using the reaction stoichiometry, resulting in a capacity of 1.0 mol
CNTs [14]. of CO2 per mole of amine under wet conditions, and 0.5 mol of CO2
Scanning electron microscopy (SEM) images, Fig. 6, show that per mole of amine under dry conditions [21]. The reactions under
the tubular morphology remained unchanged after modification wet conditions include
of CNTs with KOH. The MCNTs/TEPA image (Fig. 6c) shows a
R1 NH2 þ CO2 þ H2 O ¢ ðR1 NHþ3 ÞðHCO3 Þ ðR1Þ
homogenous morphology with swelling in the tubular structure
of MCNTs, suggesting that TEPA chains were introduced into the
R1 R2 NH þ CO2 þ H2 O ¢ ðR1 R2 NHþ2 ÞðHCO3 Þ ðR2Þ
pores of the multilayer MCNTs structure. The schematic represen-
tation of MCNTs/TEPA is illustrated in Fig. 7. It has been reported The reactions under dry conditions are
that organic solvents with low surface tension (<72 mN/m), such
as ethanol, can act as carriers for the introduction of solutes inside 2ðR1 NH2 Þ þ CO2 ¢ ðR1 NHþ3 ÞðR1 NHCOO Þ ðR3Þ
the nanotubes by capillarity [18,19]. The same process may occur
in this work by introducing TEPA inside MCNTs cavities (Fig. 7) 2ðR1 R2 NHÞ þ CO2 ¢ ðR1 R2 NHþ2 ÞðR1 R2 NCOO Þ ðR4Þ
from the use of ethanol in the preparation of the MCNTs/TEPA
sorbent. ðR1 NH2 Þ þ ðR1 R2 NHÞ þ CO2 ¢ ðR1 NHþ3 ÞðR1 R2 NCOO Þ ðR5Þ

3.4. Effect of sorption temperature on the CO2 sorption capacity

3.2. Effect of TEPA loading on the CO2 sorption capacity In addition to TEPA loading and moisture content, sorption
capacity is subject to the temperature of adsorption. To study the
In regards to CO2 sorption, the availability of amine groups in effect of adsorption temperature on CO2 capacity, a loading of
amine-based sorbents has a crucial role, as the interaction between 75 wt% MCNTs/TEPA was selected. Fig. 10 shows that with an
the CO2 and amine groups results in adsorption. Hence, the TEPA increasing temperature (25–60 °C) the sorption capacity increased.
loading amount has a direct influence on the CO2 sorption capacity. It has been proposed in other work [22,23] that the increased tem-
Fig. 8 shows that with an increasing TEPA loading (50–75 wt%) CO2 perature may be overcoming the kinetic barrier and enabling mass
sorption capacity increases from 3 to 5 mmol/g. Further increase in transfer of CO2 into the bulk of polymer. The accessibility of amine
TEPA loading (80 wt%) results in a reduction of the CO2 sorption sites increase with an increasing temperature, with the maximum
capacity. A decrease in accessible amine groups due to TEPA sorption capacity being obtained at 60 °C. The sorption rate is con-
agglomeration may describe this reduction in capacity. This parti- trolled by CO2 diffusion onto the sorbent surface at low tempera-
cle agglomeration was visible when TEPA loading was greater than tures, in this case below 60 °C. A temperature of 70 °C showed a
80 wt%, causing the sorbent to be unusable for CO2 sorption tests. decrease in capacity, this is due to adsorption being an exothermic
In addition to MCNTs, unmodified CNTs were tested as a sup- process, which causes a reduced equilibrium sorption capacity. In
port, resulting in an optimum TEPA loading of 50 wt% with a max- other words, the thermodynamics govern over the kinetics, result-
imum CO2 sorption capacity of 2.5 mmol/g. These results were ing in a decrease of CO2 sorption capacity above 60 °C [22].
obtained with a gas mixture containing 10 vol% CO2 in N2 with
1 vol% H2O. The lower surface area of unmodified CNTs, compared 3.5. Adsorption kinetics and thermodynamics
to that of MCNTs, may be attributed to the lower observed CO2
capacity. Consequently, the more amine groups loaded, the higher In this work, the Langmuir adsorption model [24] has been
CO2 sorption capacity. applied to study the adsorption kinetics of CO2 onto the adsor-
bents. Langmuir adsorption theory is based on the assumptions
that the rates of adsorption and desorption on the surface are equal
Table 1
The texture properties of CNTs, MCNTs and MCNTs/TEPA with TEPA loading of 75 wt
at equilibrium, a monolayer of the adsorbed species is formed on
%. the surface, and the rate constants of adsorption and desorption
are not functions of the number of adsorbed molecules on the sur-
Sample BET surface area (m2/g) BJH pore volume (cm3/g)
face [25,26].
CNTs 90.3 1.16 The CO2 adsorption rate on the surface is shown below [25,26]:
MCNTs 169.8 1.97
MCNTs /TEPA 2.2 0.03
14 M. Irani et al. / Fuel 206 (2017) 10–18

Fig. 5. TEM images of (a) unmodified CNTs, (b) MCNTs.

Fig. 6. SEM images of the (a) CNTs and (b) MCNTs, and (c) MCNTs/TEPA with TEPA loading of 75 wt%.

Fig. 7. Schematic illustration of TEPA loaded on MCNTs via introduction inside cavities, where black, white, and blue spheres represent C atoms, H atoms, and N atoms,
respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
M. Irani et al. / Fuel 206 (2017) 10–18 15

Fig. 10. The effect of temperature on CO2 sorption capacity at a CO2 concentration
Fig. 8. The effect of TEPA on CO2 sorption capacity at a CO2 concentration of 10 vol%
of 10 vol% with the balance as N2 gas, a gas flow rate of 300 mL/min, a TEPA loading
with the balance as N2 gas, a gas flow rate of 300 mL/min, an adsorption
of 75 wt%, 1 vol% moisture (H2O) and 50 mg sorbent.
temperature of 60 °C, 1 vol% moisture (H2O) and 50 mg of sorbent.

which describes surface coverage in terms of capacity q, and maxi-


mum capacity, qmax, results in

q ¼ qmax ½1  expðka C 0 tÞ ðE4Þ

q ¼ qmax expðkd tÞ: ðE5Þ


Adsorption measurements are applied to calculate values of ka
and kd by regressing Eqs. (4) and (5), respectively. A representative
fit for adsorption and desorption data is illustrated in Fig. 11. A
quality fit was achieved with correlation coefficients of 0.999 and
0.998 for the adsorption and desorption curves respectively. It
should be noted that the Langmuir adsorption model assumes that
the adsorbed species form a monolayer on the surface. Although
the adsorption in this work may indeed be multilayer, the Lang-
muir model was applied for our quantitative analysis because of
its simplicity.
An efficient CO2 sorbent should have fast adsorption and des-
orption kinetics at low temperatures. In this work, the adsorption
and desorption kinetics were evaluated using the Arrhenius equa-
tion described below [27]
Ea
Lnk ¼  þ LnA ðE6Þ
Fig. 9. The effect of moisture on CO2 sorption capacity at a CO2 concentration of RT
10 vol% with the balance as N2 gas, a gas flow rate of 300 mL/min, an adsorption
temperature of 60 °C, a TEPA loading of 75 wt%, and 50 mg of sorbent. which includes the reaction rate constant (k), the activation energy
(Ea), the molar gas constant (R), the reaction temperature (T), and
the frequency factor (A). The activation energy and frequency factor
are calculated by plotting Ln k versus (1/T) with a linear fit as illus-
dh trated in Fig. 12.
¼ ka C 0 ð1  hÞ ðE1Þ
dt Complete desorption of CO2 from the sorbent occurred at a low
where h is the fraction of surface coverage, t is time, ka is the temperature (90 °C) within a short time period (5 min). The
adsorption rate constant, and C0 is the concentration or partial pres- MCNTs/TEPA CO2 adsorption and desorption Ea values of 16.2 kJ/-
sure of CO2 in the gas. The CO2 desorption rate from the surface is mol and 39.9 kJ/mol, respectively, are relatively low. These results
written as show that the sorbent rate of CO2 adsorption and desorption is
high. Lower CO2 capture costs can be achieved due to this high
dh sorption rate, thus decreasing the necessary sorbent content.
¼ kd h ðE2Þ
dt Decreased sorbent content can result in a decreased operating cost
where kd is the desorption rate constant. The integration and com- because of a reduced temperature change and pressure drop in the
bination of these two equations in conjunction with the flue gas [27]. Compared to literature Ea values of aqueous amine-
relationship based absorption (40–50 kJ/mol), reported solid adsorption Ea val-
ues are generally lower [16,28]. The activation energy for desorp-
q ¼ qmax h ðE3Þ tion in this work is comparable to the results of other solid
16 M. Irani et al. / Fuel 206 (2017) 10–18

Fig. 11. The CO2 adsorption (a) and the CO2 desorption (b) curves MCNT/TEPA (blue lines: experimental data, red lines: Langmuir model fit to obtain ka and kd) [CO2
concentration: 10 vol%; balance gas: N2; gas flow rate: 300 mL/min; H2O: 1 vol%; adsorption temperature: 60 °C; TEPA loading:75 wt%; desorption temperature: 90 °C;
desorption carrier gas: N2 with a flow rate of 300 mL/min; weight of sorbent: 50 mg].

Fig. 12. Arrhenius plots of MCNT/TEPA for adsorption (a) and desorption (b). [CO2 concentration: 10 vol%; balance gas: N2; gas flow rate: 300 mL/min; H2O: 1 vol%; TEPA
loading: 75 wt%; desorption carrier gas: N2 with a flow rate of 300 mL/min; weight of sorbent: 50 mg].

amine-based sorbents, for example, Liu et al. [16] prepared a sor- tion energy for desorption was in the range of 41–46 kJ/mol, with a
bent using impregnation of tetraethylenepentamine (TEPA) in CO2 capacity of around 4 mmol/g using 2 vol% CO2 gas at 16 °C.
industrial grade multiwalled carbon nanotubes (IG-MWCNTs) Therefore, MCNTs/TEPA has potential to be an efficient CO2 adsor-
and calculated the activation energy of CO2 adsorption/desorption bent because of its low energy requirement and its high CO2 sorp-
using the Arrhenius equation. Their results showed that the activa- tion capacity.
M. Irani et al. / Fuel 206 (2017) 10–18 17

Acknowledgment

Authors highly appreciate the support of the state of Wyoming.


They would like to thank Sabrina Kaufman, Kyle Summerfield and
Emma Jane Alexander from Shell 3D Visualization Resources at
School of Energy and Resources. Authors would like also thank to
Joshua Stecher for TEM images.

References

[1] Mondal MK, Balsora HK, Varshney P. Progress and trends in CO2 capture/
separation technologies: a review. Energy 2012;46:431–41.
[2] Goeppert A, Czaun M, Prakash GS, Olah GA. Air as the renewable carbon source
of the future: an overview of CO2 capture from the atmosphere. Energy Environ
Sci 2012;5:7833–53.
[3] Leung DY, Caramanna G, Maroto-Valer MM. An overview of current status of
carbon dioxide capture and storage technologies. Renewable Sustainable
Energy Rev 2014;39:426–43.
[4] Aaron D, Tsouris C. Separation of CO2 from flue gas: a review. Sep Sci Technol
Fig. 13. Cyclic CO2 sorption and desorption [sorption conditions: CO2 concentra- 2005;40:321–48.
tion, 10 vol%; gas flow rate, 300 mL/min; H2O: 1 vol%; sorption temperature, 60 °C; [5] Chandra V, Yu SU, Kim SH, Yoon YS, Kim DY, Kwon AH, et al. Highly selective
TEPA loading: 75 wt%; desorption conditions:N2 flow rate, 300 mL/min; desorption CO2 capture on N-doped carbon produced by chemical activation of
temperature, 90 °C; weight of sorbent:50 mg]. polypyrrole functionalized graphene sheets. Chem Commun 2012;48:735–7.
[6] Dutcher B, Fan M, Russell AG. Amine-based CO2 capture technology
development from the beginning of 2013 a review. ACS Appl Mater
Interfaces 2015;7:2137–48.
Furthermore, the heat of the reaction can be studied as follows [7] Kenarsari SD, Yang D, Jiang G, Zhang S, Wang J, Russell AG, et al. Review of
[27] recent advances in carbon dioxide separation and capture. Rsc Adv
2013;3:22739–73.
[8] Samanta A, Zhao A, Shimizu GK, Sarkar P, Gupta R. Post-combustion CO2
DH ¼ Ea;ads  Ea;des ðE7Þ capture using solid sorbents: a review. Ind Eng Chem Res 2011;51:1438–63.
[9] Grondein A, Bélanger D. Chemical modification of carbon powders with
aminophenyl and aryl-aliphatic amine groups by reduction of in situ generated
where DH is the heat of the reaction, and Ea;ads and Ea;des are the acti- diazonium cations: applicability of the grafted powder towards CO2 capture.
vation energies of adsorption and desorption, respectively, obtained Fuel 2011;90:2684–93.
from the Arrhenius equation (E6). DH was calculated to be [10] Ye Q, Jiang J, Wang C, Liu Y, Pan H, Shi Y. Adsorption of low-concentration
carbon dioxide on amine-modified carbon nanotubes at ambient temperature.
23.7 kJ/mol for MCNTs/TEPA. The sorption of CO2 onto MCNTs/ Energy Fuels 2012;26:2497–504.
TEPA is confirmed to be exothermic by the negative value of DH. [11] Liu Y, Ye Q, Shen M, Shi J, Chen J, Pan H, et al. Carbon dioxide capture by
functionalized solid amine sorbents with simulated flue gas conditions.
Environ Sci Technol 2011;45:5710–6.
[12] Smart S, Cassady A, Lu G, Martin D. The biocompatibility of carbon nanotubes.
3.6. Regenerability Carbon 2006;44:1034–47.
[13] Su F, Lu C, Cnen W, Bai H, Hwang JF. Capture of CO2 from flue gas via
In an industrial settings, an effective sorbent should have not multiwalled carbon nanotubes. Sci Total Environ 2009;407:3017–23.
[14] Raymundo-Pinero E, Azais P, Cacciaguerra T, Cazorla-Amorós D, Linares-
only high sorption capacity but also needs to be regenerable Solano A, Béguin F. KOH and NaOH activation mechanisms of multiwalled
through several sorption–desorption cycles. In this work, the study carbon nanotubes with different structural organisation. Carbon
of sorbent regenerability was performed in 10 sorption-desorption 2005;43:786–95.
[15] Liu Q, Shi Y, Zheng S, Ning L, Ye Q, Tao M, et al. Amine-functionalized low-cost
cycles. The results from Fig. 13 show that the sorption capacity
industrial grade multi-walled carbon nanotubes for the capture of carbon
reduced with an increase in cycle number, with a reduction of dioxide. J Energy Chem 2014;23:111–8.
about 20% (5 mmol/g to 4 mmol/g) after 10 cycles. TEPA degrada- [16] Liu Q, Shi J, Zheng S, Tao M, He Y, Shi Y. Kinetics studies of CO2 adsorption/
desorption on amine-functionalized multiwalled carbon nanotubes. Ind Eng
tion [8] and evaporation, most likely due to the low molecular
Chem Res 2014;53:11677–83.
weight of TEPA, may be attributed to the loss of CO2 sorption [17] Zhu B-K, Xie S-H, Xu Z-K, Xu Y-Y. Preparation and properties of the
capacity. It is suggested that using amine functional components polyimide/multi-walled carbon nanotubes (MWNTs) nanocomposites.
with higher molecular weights than that of TEPA may give greater Compos Sci Technol 2006;66:548–54.
[18] Ebbesen TW. Wetting, filling and decorating carbon nanotubes. J Phys Chem
thermal stability, lessening the likelihood of evaporation and Solids 1996;57:951–5.
decomposition during sorption–desorption cycles, resulting in bet- [19] Tsang S, Chen Y, Harris P, Green M. A simple chemical method of opening and
ter regenerability of the sorbent. filling carbon nanotubes. Nature 1994;372:159–62.
[20] Mohammad SA, Gasem KAM. Multiphase analysis for high-pressure
adsorption of CO2/water mixtures on wet coals. Energy Fuels
2012;26:3470–80.
4. Conclusions [21] He L, Fan M, Dutcher B, Cui S, Shen X-D, Kong Y, et al. Dynamic separation of
ultradilute CO2 with a nanoporous amine-based sorbent. Chem Eng J
2012;189–190:13–23.
A sorbent for CO2 capture has been developed by loading TEPA [22] Wang W, Xiao J, Wei X, Ding J, Wang X, Song C. Development of a new clay
on modified carbon nanotubes (MCNTs). An optimal TEPA loading supported polyethylenimine composite for CO2 capture. Appl Energy
2014;113:334–41.
of 75 wt% obtained the maximal CO2 sorption capacity of 5 mmol/g [23] Ma X, Wang X, Song C. ‘‘Molecular basket” sorbents for separation of CO2 and
for a gas mixture containing 10 vol% CO2 in N2 with 1 vol% H2O. H2S from various gas streams. J Am Chem Soc 2009;131:5777–83.
Unmodified CNTs as a support showed a maximal CO2 sorption [24] Ko YG, Shin SS, Choi US. Primary, secondary, and tertiary amines for CO2
capture: designing for mesoporous CO2 adsorbents. J Colloid Interface Sci
capacity of 2.5 mmol/g with an optimal TEPA loading of 50 wt%.
2011;361:594–602.
Thermal stability studies indicate that MCNTs/TEPA is stable below [25] Chen M-S, Fan H-F, Lin K-C. Kinetic and thermodynamic investigation of
110 °C. MCNTs/TEPA kinetic data of CO2 adsorption and desorption Rhodamine B adsorption at solid/solvent interfaces by use of evanescent-wave
show a low energy requirement and a short-time cycle for regen- cavity ring-down spectroscopy. Anal Chem 2009;82:868–77.
[26] Zhou X, Yi H, Tang X, Deng H, Liu H. Thermodynamics for the adsorption of
eration; therefore, MCNTs/TEPA can greatly decrease the overall SO2, NO and CO2 from flue gas on activated carbon fiber. Chem Eng J
cost of capturing CO2. 2012;200:399–404.
18 M. Irani et al. / Fuel 206 (2017) 10–18

[27] Dutcher B, Fan M, Leonard B, Dyar MD, Tang J, Speicher EA, et al. Use of [28] Paul S, Ghoshal AK, Mandal B. Absorption of carbon dioxide into aqueous
nanoporous FeOOH as a catalytic support for NaHCO3 decomposition aimed at solutions of 2-Piperidineethanol: kinetics analysis. Ind Eng Chem Res
reduction of energy requirement of Na2CO3/NaHCO3 based CO2 separation 2009;48:1414–9.
technology. J Phys Chem C 2011;115:15532–44.

Das könnte Ihnen auch gefallen