You are on page 1of 9

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/290435394

Linear Driving Force Model in Carbon Dioxide Capture by Adsorption

Article  in  Applied Mechanics and Materials · March 2016


DOI: 10.4028/www.scientific.net/AMM.830.38

CITATIONS READS

2 1,339

7 authors, including:

Leonardo Hadlich De Oliveira Edson Antonio da Silva


Universidade Estadual de Maringá Universidade Estadual do Oeste do Paraná
31 PUBLICATIONS   246 CITATIONS    203 PUBLICATIONS   2,394 CITATIONS   

SEE PROFILE SEE PROFILE

Maria Angélica S. D. Barros Pedro Augusto Arroyo


Universidade Estadual de Maringá Universidade Estadual de Maringá
71 PUBLICATIONS   779 CITATIONS    136 PUBLICATIONS   1,138 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

H2S and CO2 Pollutants Removal from Pre-salt Natural Gas by Adsorption Processes View project

Subprodutos industriais View project

All content following this page was uploaded by Leonardo Hadlich De Oliveira on 12 April 2016.

The user has requested enhancement of the downloaded file.


LINEAR DRIVING FORCE MODEL IN CARBON DIOXIDE CAPTURE BY
ADSORPTION
Leonardo Hadlich de Oliveira1,a, Joziane Gimenes Meneguin2,b, Edson
Antonio da Silva3,c, Maria Angélica Simões Dornellas de Barros4,d, Pedro
Augusto Arroyo5,e,*, Wilson Mantovani Grava6,f, Jailton Ferreira do
Nascimento7,g
1
Laboratory of Adsorption and Ion Exchange, Chemical Engineering Department, State University
of Maringá, Av. Colombo 5790, Bl. E10, 87020-900, Maringá – PR – Brazil
2
Laboratory of Adsorption and Ion Exchange, Chemical Engineering Department, State University
of Maringá, Av. Colombo 5790, Bl. E10, 87020-900, Maringá – PR – Brazil
3
Chemical Engineering Department, State University of West Paraná, R. Faculdade 645, 85903-
000, Toledo – PR – Brazil
4
Laboratory of Adsorption and Ion Exchange, Chemical Engineering Department, State University
of Maringá, Av. Colombo 5790, Bl. E10, 87020-900, Maringá – PR – Brazil
5
Laboratory of Adsorption and Ion Exchange, Chemical Engineering Department, State University
of Maringá, Av. Colombo 5790, Bl. E10, 87020-900, Maringá – PR – Brazil
6
PETROBRAS S.A., CENPES R&D Centre/PDEP/PRO-CO2, Av. Horácio Macedo, 950, 21941-
915, Rio de Janeiro – RJ – Brazil
7
PETROBRAS S.A., CENPES R&D Centre/PDEP/TPP, Av. Horácio Macedo, 950, 21941-915, Rio
de Janeiro – RJ – Brazil
a
leonardoh.deoliveira@gmail.com, bjozigimenes@gmail.com, cedsondeq@hotmail.com,
d
angelicabarros.deq@gmail.com, earroyo@deq.uem.br, fwilson.grava@petrobras.com.br,
g
jfer@petrobras.com.br
* corresponding author

Keywords: CO2 capture, NaY zeolite, LDF model.

Abstract. Modelling of CO2 capture by adsorption is an important subject for the industry of
separation technology. It is used, for example, in studies of CO2 separation from flue gases and CO2
removal from natural gas. In this work, experimental data of CO2 capture by adsorption was
determined gravimetrically, at 30 °C and pressures up to 40 bar, and in a dynamic adsorption unit,
at 30°C and 20 bar, using a NaY zeolite fixed bed column. The Langmuir model was used to
correlate the experimental equilibrium data and presented a maximum adsorbed amount equal to
7.05 mmol/g, which represents a high adsorption capacity. The breakthrough experimental data
were modelled using Linear Driving Force (LDF) and Thomas models. The lumped mass transfer
coefficient Ks is only adjustable parameter of LDF model, and this model represented better the CO2
breakthrough curve than Thomas model. The LDF model presented a good correlation with
experimental data with a sum of square deviation equal to 0.2081. The Ks coefficient was estimated
as 0.0872 min-1, which is smaller than the overall volumetric mass transfer coefficient (kca = 0.9879
min-1), obtained experimentally, indicating that the intraparticle resistance is higher than external
film resistance. Therefore, the behavior of CO2 breakthrough curve can be represented by LDF
model.

Introduction

The most rigorous mathematical models used to represent the dynamics of gas adsorption in
fixed bed columns are composed by partial differential equations, which have three independent
variables: time, bed space coordinate and adsorbent particle space coordinate in axial direction.
Moreover, equations that represent solid-gas interface equilibrium are also necessary. Therefore, in
order to reduce the complexity of the modelling, concentrated parameters model can be employed
to particle or adsorbed phase. In this context, the Linear Driving Force (LDF) was utilized. LDF
model was originally proposed by Gleuckauf and Coates [1]. It is considered a practical tool to
study the adsorption kinetics on heterogeneous solids, evaluate adsorbent fixed bed dynamics,
correlate and predict mono and multicomponent gas adsorption, and to design adsorptive processes
as well [2]. Linear driving force model was much investigated for gas separation in 80’s decade [3]
and has been used to describe the dynamic and batch adsorption behavior, using different
adsorbents and flow conditions [4,5,6,7,8,9,10,11].
Although being of great importance in natural gas separation, the CO2 uptake at high
pressures has been scarcely investigated, mainly when understanding the adsorption mechanism and
modeling are the main objectives. Therefore, this work aimed to obtain CO2 uptake in a NaY
packed bed at 30 °C and 20 bar. To reach this goal, characterization of the material, equilibrium
isotherm of CO2 adsorption, and calculation of mass transfer parameters were also determined, and
the LDF model was used to correlate experimental data and compared to the classical Thomas
model [12].

Experimental

Characterization. The Si/Al molar ratio of NaY zeolite was determined by a Varian SpectrAA 50B
atomic absorption spectrometer. DRX analysis was carried out in a Shimadzu XRD-6000 X-ray
diffractometer using Cu-Kα monochromatic radiation (λ = 1.5418 Å, 5 < 2θ < 80 °, step = 0.01 °,
and scan speed = 1 °/s). Physisorption of N2 was determined in a Micromeritics ASAP 2020
analyzer at -196 °C, using pure gas in a relative pressure 0.0001 ≤ P/P0 ≤ 1, after sample outgassing
at 5 μmHg, 150 °C for 10 h. The textural properties of the zeolitic sample were estimated using
BET and t-plot [13].

Adsorption equilibrium data. A known amount of NaY was inserted in the sample container of a
Rubotherm magnetic suspension balance (MSB), precise to ± 1×10-6 g. The sample was activated at
150 °C for 10 h under vacuum. The adsorption step was determined injecting CO2 in the system and
estimating the adsorbed mass for 0.4 up to 40 bar, at 30 °C, and applying the mass balance
presented by Dreisbach et al. [14]. The amount of CO2 adsorbed was determined by applying a
mathematical treatment similar to Keller and Staudt [15] and Murata et al. [16], considering that for
densities greater than a critical value the adsorbent saturates.

Fixed bed column adsorption. The dynamic adsorption column consisted of stainless steel tube 2
cm internal diameter and 30 cm long and contained 26.96 g of NaY with average particle size of
1.44 mm resulting in a packed 22 cm of packed bed. The run was performed at 30 °C and 20 bar,
using an inlet flow rate of 7.1 mLCO2/min. The composition of the outlet stream was analyzed with
a gas chromatograph (Bruker 450 GC).

Determination of mass transfer parameters. From experimental breakthrough curve, the mass
transfer zone (MTZ), overall volumetric coefficient (Kca), and operational ratio (Ro) as proposed by
Barros et al. [17]. The amount of CO2 adsorbed at breakthrough and saturation times, qtu and qtt,
was estimated through a mass balance.

Modelling

Modelling of equilibrium data. Adsorption equilibrium data were correlated with Langmuir
model. The parameters were fitted using Evolutionary optimization method available in Microsoft
Excel®. The objective function (Feq) was:

𝑒𝑥𝑝 2
𝐹𝑒𝑞 = ∑𝑛𝑖=1(𝑞𝑒𝑞 𝑐𝑎𝑙𝑐
− 𝑞𝑒𝑞 ) . (1)

Where: n is the number of data set, qeq is the equilibrium adsorbed amount of CO2, and superscripts
exp and calc refer to the experimental and calculated, respectively.

Modelling of the adsorption in fixed bed column. The modelling employed the following
assumptions: plug flow pattern (no axial dispersion); no radial distribution of gas concentration;
isothermic bed operation; negligible pressure drop.
The mass balance for gas phase, for a given bed volume element, can be describe as:

𝜕𝐶(𝑡,𝑧) 𝜕𝐶(𝑡,𝑧) 𝜌𝑏𝑒𝑑 𝜕𝑞(𝑡,𝑧) (2)


+ 𝑢𝜀 + = 0.
𝜕𝑡 𝜕𝑧 𝜀 𝜕𝑡

Where: C is the gas phase concentration (mmol/L); q is the gas adsorbed quantity (mmol/g); uε is
the interstitial velocity (cm/min); ρbed is the bed density (g/L); t is the time (min); z the distance
along the bed (cm); and ε is the bed porosity, estimated using the equation correlation proposed by
Ribeiro et al. [18]:

𝑑𝑖𝑛𝑡 (3)
𝜀 = 0.373 + 0.917exp⁡(−0.824 ).
𝑑𝑃

Where: dint and dp are the internal diameter of the column (cm) and the particle diameter (mm),
respectively.
The LDF model represents the gas mass transfer in adsorbent phase as:

𝜕𝑞(𝑡,𝑧) (4)
= −𝐾𝑆 (𝑞(𝑡, 𝑧) − 𝑞𝑒𝑞 ).
𝜕𝑡

Where KS is the lumped mass transfer coefficient (min-1).


In Thomas model, the adsorption rate is represented by the Langmuir kinetics [12]:

𝜕𝑞(𝑡,𝑧)
= 𝑘𝑎 (𝑞𝑚𝑎𝑥 − 𝑞(𝑡, 𝑧))𝐶 − 𝑘𝑑 𝑞(𝑡, 𝑧). (5)
𝜕𝑡

Where ka (L/mmol min), kd (min-1) are the adsorption and desorption rates, respectively, and qmax is
the maximum adsorbed capacity (mmol/g).
To solve both models, the initial and boundary conditions used were:

𝑞(𝑡, 𝑧) = 0, for t = 0. (6)

𝐶(𝑡, 𝑧) = 0, for t = 0. (7)

𝐶(𝑡, 𝑧) = 𝐶0 , for z = 0. (8)

Where: C0 is the inlet gas phase concentration (mmol/L).


The line method [19] was used to solve the differential equation system with a Fortran
computational program. The subroutine DASSL, developed by Petzold [20], was utilized in the
solution ODE system.
In LDF model, Ks is the only adjustable parameter, and was determined using the Golden
Section Search optimization method [21]. Thomas model has ka and kd as adjustable parameters,
which were estimated using the downhill simplex method [22] optimization method.
For both models the objective function F used was:
2
𝐶 𝑒𝑥𝑝 𝐶 𝑐𝑎𝑙𝑐
𝐹 = ∑𝑛𝑖=1 [(𝐶 ) − (𝐶 ) ] . (9)
0 0

The experimental input variables used in modeling calculations were: ρbed = 390.04 g/L; ε =
0.373; uε = 6.66 cm/min; Q = 0.78×10-2 L/min; and finally C0 = 892 mmol/L.

Results and discussion

According to atomic mass absorption spectrometry, the Si/Al molar ratio was 2.8, with a cell
composition of Na51[(AlO2-)51 . (SiO2)141], which is typical of synthetic NaY zeolite [23]. Such
result is in agreement with the most intense peaks at 2θ = 6.3, 10.3, 15.9, and 23.2º, shown in XRD
pattern (Fig. 1a), which also indicates the high crystallinity of FAU framework with cell parameters
ao = 24 Å [24,25]. N2 physisorption presented in Fig. 1b is classified as type I, commonly found in
microporous materials, which is the case of zeolites [26]. Indeed, the XRD pattern is attributed to a
NaY framework.
10000 270

 260
8000

250
Q (cm³/g STP)
I (Counts.)

6000
240

 
4000 230

220
2000 
210

10 20 30 40 50 60 200
0.0 0.2 0.4 0.6 0.8 1.0
(º)
P/P0

(a) (b)
Figure 1. Characterization of NaY zeolite. a) XRD pattern; b) N2 sorption isotherm.

The BET area was 672 m²/g. The total pore volume and the microporous volume were
estimated as 0.405 and 0.322 cm3/g, respectively. As the volume ratio was 0.94, the micropores
predominance was confirmed.
Adsorption equilibrium data for CO2 in NaY at 30 °C and the Langmuir fit are shown in Fig.
2. The fitted parameters were qmax = 7.05 ± 0.03 mmol/g and b = 0.050 ± 0.003 L/mmol. The small
objective function value F of 1.568 as well as the parameter deviations made us conclude that this
model may successfully represent the adsorption of CO2 in NaY zeolite.

Figure 2. Adsorption isotherm of CO2 in NaY zeolite at 30 °C. Experimental data (♦) and calculated
using Langmuir model (line).
The Langmuir model is referred to pseudo-Langmuir, since it indicates the saturation of the
adsorbent similar to a monolayer fully covered. This behaviour was already expected since NaY
zeolite is a strong adsorbent for CO2 with an almost irreversible isotherm [27].
It can be seen that the maximum adsorbed amount of CO2 on the zeolite sample was close to
7 mmol/g, which is in agreement with previous results for faujasites [28].
However, 70% of this amount was achieved at 1 bar, indicating a fast saturation of the
porous structure. Such phenomenon was a consequence of a strong interaction between CO2
molecules and the surface. In fact, the high-energy sites are first occupied by CO2 molecules
[29,30,31,32], mainly in the ones of NaY as it has high adsorption capacity [28].
The Langmuir model assumes that adsorption occurs in specific sites of the surface with
equivalent energy. Therefore, the Langmuir equation is based on the assumption of a structurally
homogeneous adsorbent [30]. In addition, the larger b, the stronger the affinity of the adsorbate
molecule towards the surface.
In summary, the mechanism proposed for CO2 adsorption in zeolite NaY indicates that this
process mostly occurs in the cavities, since it is a predominant microporous material, as discussed
earlier. Probably, linear OCO–X+ complexes were formed (X+ is the cation) involving the
perturbation of Si–O–Al bonds [30,33,34].
Experimental breakthrough curve is shown in Fig. 3. From these data, the mass transfer
parameters (Table 1) were calculated.

Figure 3. Experimental breakthrough curve for CO2 adsorption in NaY zeolite fixed bed column at
30 °C and 20 bar.

Table 1. Mass Transfer Parameters.


tb 16.16 [min]
ts 60.81 [min]
t* 26.7 [min]
MTZ 10.11 [cm]
Kca 0.9879 [min-1]
Ro 0.87

From results in Fig. 3 and Table 1 it was verified that breakthrough time tb is smaller than
the ideal adsorption time t*, indicating experimental consistency. The size of Area 1 (2.42 min) was
smaller than Area 2 (5.44 min). It means the existence of important mass transfer resistances related
not only to film but also to intraparticle diffusion, related to the high microporosity of the adsorbent
[35].
Parameter Kca was calculated considering irreversible isotherm [35]. Since CO2 isotherm
behavior (Fig. 2) was close to an irreversible one, results shown in Fig. 4 were determined from
experimental breakthrough data, calculating N(τ - 1) = ln(C/C0) – 1. Such expression was obtained
from an analytical solution of mass balance in the bed. The shape of Fig. 4 was related to external
mass transfer resistances.
Figure 4. Breakthrough curve for irreversible adsorption.

The mass transfer parameter Ro was equal to 0.87, which indicates that the column was not
operating near optimal condition [36], that is a step shape. An optimization should be carried out in
order to achieve Ro close to 0 in order to minimize the external mass transfer resistances in the
packed bed.
Furthermore, the amount of CO2 adsorbed up to breakpoint time (qtu) and saturation time
(qtt) were estimated 4.14 mmol/g and 7.66 mmol/g, respectively. It must be emphasized that qtt was
very similar to qeq (deviation less than 10%), which indicates that the system reached
thermodynamic equilibrium.
Experimental CO2 breakthrough data and model adjustments are shown in Fig. 5.

1.0

0.8

0.6
C/C0
0.4

0.2

0.0
0 20 40 60 80
t (min)
Figure 5. Breakthrough curve for CO2 adsorption in NaY zeolite fixed bed column at 30 °C and 20
bar. Experimental (solid line) and calculated with LDF (dashed line) and Thomas (dotted) models.

Both models estimated CO2 saturation time in agreement with experimental data.
Nevertheless, LDF model better represented the breakpoint time as well as the whole curve. LDF
parameter obtained was KS = 0.0872 min-1, with F = 0.2081. The parameters estimated for Thomas
model were ka = 2.0318·10-4 L/(mmol min) and kd = 1.2743·10-2 min-1, with a higher F = 0.4113.
As LDF takes into account a predominant intraparticle resistance, the intraparticle diffusion should
be also considered. Indeed, the experimental breakthrough curve behaviour showed a non-
symmetric shape (Area 2 > Area 1 in Fig. 3), indicating the diffusional resistance mainly in the
zeolite channels and cavities. Such result was already expected due to microporous characteristic of
the zeolitic sample. Moreover, DRX results show a high crystalline zeolite without amorphous
impurity blocking the pore system.
In this way, LDF model represented the experimental breakthrough curve despite of
assumption of intraparticle resistance as the controlling step, although mass transfer parameters also
indicated the importance of film resistances. It may be concluded that both resistances should be
considered, and intraparticle resistance may be higher than external film resistance, as KS < kca in
the investigation presented herein.

Conclusions

According to results presented herein, it is possible to assure that NaY zeolite has a strong
affinity to CO2 molecules, as the isotherm is almost irreversible. The maximum adsorbed amount,
estimated using Langmuir model, was estimated as 7.05 mmol/g, showing that the model fitted
adequately the experimental isotherm. Indeed, even in not optimized conditions, the breakthrough
experimental data lead to an adsorption capacity at saturation with a deviation less than 10% when
compared to equilibrium data. LDF and Thomas models were compared in fitting experimental
breakthrough curve, and LDF model can be successfully used to correlate CO2 uptake in the NaY
zeolite fixed bed column. The model presented a small sum square deviation between experimental
and calculated of 0.2081. The lumped mass transfer coefficient Ks, used as adjustable parameter,
was equal to 0.0872 min-1. Since this value is smaller than the overall volumetric mass transfer
coefficient (kca = 0,9879 min-1), the intraparticle resistance is assumed higher than external film
resistance. Finally, the mass transfer resistances in NaY micropores was assumed as a limiting step
for the system studied although external film resistance should be considered.

Acknowledgements

The authors would like to thank the financial support of Petrobras.

References

[1] E. Gleuckauf, J.I. Coates, Theory of Chromatography. Part IV. The Influence of Incomplete
Equilibrium on the Front Boundary of Chromatograms and the Effectiveness of Separation, J.
Chem. Soc. 0 (1947) 1315-1321.
[2] S. Sircar, J.R. Hufton, Why Does the Linear Driving Force Model for Adsorption
KineticsWork? Adsorption 6 (2000) 137-147.
[3] R.T. Yang, S.J. Doong, Bulk separation of multicomponent gas mixtures by pressure swing
adsorption: pore/surface diffusion and equilibrium models, AIChE J. 32 (1986) 397-410.
[4] K. Chihara, M. Suzuki, Simulation of nonisothermal pressure swing adsorption, J. Chem. Eng.
Japan 16 (1983) 53-61.
[5] S. Sircar, Linear-driving-force Model for Non-isothermal Gas Adsorption Kinetics, J. Chem.
Soc., Faraday Trans. I 79 (1983) 785-796.
[6] K.S Hwang, J.H. Jun, W.K. Lee, Fixed-bed adsorption for bulk component system. Non-
equilibrium, non-isothermal and non-adiabatic model, Chem. Eng. Sci. 50 (1995) 813-825.
[7] A.I. Fatehi, K.F. Loughlin, M.M. Hassan, Separation of methane –nitrogen mixtures by pressure
swing adsorption using a carbon molecular sieve, Gas. Sep. Purif. 9 (1995) 199-204.
[8] J.-G. Jee, S.-J. Lee, C.-H. Lee, Comparison of the adsorption dynamics of air on zeolite 5A and
carbon molecular sieve beds, Korean J. Chem. Eng., 21 (2004) 1183-1192.
[9] G.K. Prasad, B. Singh, A. Saxena, Kinetics of adsorption of sulfur mustard vapor on carbons
under static conditions, AIChE J. 53 (2006) 678-682.
[10] E. Glueckauf, Theory of Chromatography. Part 10. Formulae for Diffusion into Spheres and
Their Application to Chromatography, Trans. Faraday Soc. 51 (1995) 1540-1551.
[11] D.D. Do, R.G. Rice, Validity of the Parabolic Profile Assumption in Adsorber Studies, AIChE
J. 32 (1986) 149-154.
[12] H.C. Thomas, Heterogeneous ion exchange in a flowing system. J. Am. Chem. Soc. 66, p.
1664–1666, 1944.
[13] S,J. Gregg, K.S.W. Sing, Adsorption, Surface Area and Porosity, Academic Press, London,
1982.
[14] F. Dreisbach, H.W. Lösch, P. Harting, Highest Pressure Adsorption Equilibria Data:
Measurement with Magnetic Suspension Balance and Analysis with a New Adsorbent/Adsorbate-
Volume, Adsorption 8 (2002) 95-109.
[15] J. Keller, R. Staudt, Gas Adsorption Equilibria, Springer US, 2005.
[16] K. Murata, J. Miyawaki, K. Kaneko, A simple determination method of the absolute adsorbed
amount for high pressure gas adsorption, Carbon 40 (2002), 425–428.
[17] M.A.S.D. Barros, P.A. Arroyo, E.A. Silva, General Aspects of Aqueous Sorption Process in
Fixed Beds, in: H. Nakajima (Ed.), Mass Transfer - Advances in Sustainable Energy and
Environment Oriented Numerical Modeling, InTech, Rijeka, 2013, 361-386.
[18] A.M. Ribeiro, P. Neto, C. Pinho, Mean Porosity and Pressure Drop Measurements in Packed
Beds of Monosized Spheres: Side Wall Effects, Int. Rev. Chem. Eng. 2 (2010) 40-46.
[19] S. Pamuk, A. Erdem, The method of lines for the numerical solution of a mathematical model
for capillary formation: The role of endothelial cells in the capillary. Appl. Math. Comput. 186
(2007) 831-835.
[20] L.R. Petzold, A description of DASSL: A differential/algebraic equation system solver, STR,
SAND82-8637, Livermore 1982.
[21] T.F. Edgar, D.M. Himmelblau, Optimization of chemical process, McGraw-Hill, 1989.
[22] J.A. Nelder, R.A. Mead, Simplex method for function minimization, Computer J. 7 (1965)
308–313.
[23] D.W. Breck, Zeolite Molecular Sieves, Robert E. Krieger Publishing Company, 1974.
[24] F. Fotovat, H. Kazemian, M. Kazemeini, Synthesis of Na-A and faujasitic zeolites from high
silicon fly ash, Materials Research Bulletin 44 (2009) 913–917.
[25] M.M.J. Treacy, J.B. Higgins, Collection of Simulated XRD Powder Patterns for Zeolites, 4th
Revised Edition, Published on behalf of the Structure Commision of the International Zeolite
Association, Elsevier, 2001.
[26] IUPAC; Physical Chemistry Division, Commission on Colloid and Surface Chemistry
including Catalysis, Reporting physisorption data for gas/solid systems with special reference to the
determination of surface area and porosity, Pure Applied Chemistry 57 (1985) 603–619.
[27] W. Gao, D. Butler, D.L. Tomasko, High-Pressure Adsorption of CO2 on NaY Zeolite and
Model Prediction of Adsorption Isotherms. Langmuir 20 (2004) 8083–8089.
[28] J. Zhang, N. Burke, S. Zhang, K. Liu, M. Pryukhina, Thermodynamic analysis of molecular
simulations of CO2 and CH4 adsorption in FAU zeolites, Chem. Eng. Sci. 113 (2014) 54–61.
[29] S. Coluccia, L. Marchese, L., G. Martra, Characterisation of microporous and mesoporous
materials by the adsorption ofmolecular probes: FTIR and UV-vis studies, Microporous
Mesoporous Mater. 30 (1999) 43–56.
[30] K.S. Walton, M. B., Abney, M.D. LeVan, CO2 adsorption in Y and X zeolites modified by
alkali metal cation exchange, Microporous Mesoporous Mater. 91 (2006) 78–84.
[31] Uibu, M., Uus, M., Kuusik, R. CO2 mineral sequestration in oil-shale wastes from Estonian
power production, J. Environ. Manage. 90 (2009) 1253–1260.
[32] M. Olivares-Marín, S. García, C. Pevida, M.S. Wong, M. Maroto-Valer, The influence of the
precursor and synthesis method on the CO2 capture capacity of carpet waste-based sorbents, J.
Environ. Manage. 92 (2011) 2810-2817.
[33] G. Martra, S. Coluccia, P. Davit, E. Gianotti, L. Marchese, H. Tsuji, H. Hattori, Acidic and
basic sites in NaX and NaY faujasites investigated by NH3, CO2 and CO molecular probes, Res.
Chem. Intermed. 25 (1999) 77-93.
[34] T. Montanari, G. Busca, On the mechanism of adsorption and separation of CO2 on LTA
zeolites: an IR investigation, Vib. Spectrosc. 46 (2008) 45-51.
[35] W.L. McCabe, J.C. Smith, P. Harriot, Unit Operations of Chemical Engineering, McGraw-
Hill, Inc., 1993.
[36] R. Lambrecht, M.A.S.D. Barros, E.S. Cossich, E.A. Silva, G.K.L. Matta, R. Stachiw,
Adsorption of Reactive Blue 5G Dye by Activated Carbon and Pyrolyzed Shale Oil Residue,
Adsorp. Sci. Technol. 25 (2007) 741-749.

View publication stats

You might also like