Vous êtes sur la page 1sur 6

J. Chem.

Thermodynamics 103 (2016) 16

Contents lists available at ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Amino-functionalized ionic liquids as carbon dioxide scavengers. Ab


initio thermodynamics for chemisorption
Nadezhda A. Andreeva a, Vitaly V. Chaban b,
a
b

PRAMO, St. Petersburg, Russian Federation


Federal University of So Paulo, So Paulo, Brazil

a r t i c l e

i n f o

Article history:
Received 6 June 2016
Received in revised form 22 July 2016
Accepted 30 July 2016
Available online 1 August 2016
Keywords:
Ionic liquids
Tetraethylammonium
Monoethanolamine
Imidazolium
Pyridinium
Morpholinium
Pyrrolidinium
Pyrazolium
Thermodynamics
Infrared spectra

a b s t r a c t
Carbon dioxide (CO2) capture by aqueous amine solutions is an industry grade technology nowadays.
Some drawbacks of this technology motivate development of new CO2 scavengers. Due to low volatility,
thermal stability, non-flammability, and tunability, room-temperature ionic liquids (RTILs) are viewed as
prospective universal solvents. The amino groups can be grafted to the organic cations yielding species
with the increased CO2 capture performance. Many favorable physical properties of the original RTILs
can be retained in this way. We report systematic computational analysis (enthalpy, entropy, Gibbs free
energy, geometric parameters, partial charges, vibrational frequencies) of six organic cations, each representing a different family of RTILs. We found that the longer side hydrocarbon chain of the cations offer
the most prospective sites in view of the CO2 capture. In turn, chemisorption in the rings is less thermodynamically favorable. It was possible to corroborate the thermodynamic regularities in terms of electrophile + nucleophile organic reactions involving partial charges and structure perturbations due to
carbamate formation. The reported results foster development of alternative CO2 scavengers by eliminating their volatility and increasing reaction yields.
2016 Elsevier Ltd.

1. Introduction
Carbon dioxide (CO2) belongs to the greenhouse gases, which
are responsible for global warming. During a couple of last decades,
development of efficient CO2 removal methods became an important problem [14]. A variety of techniques, based on absorption,
adsorption, membranes, and hybrid setups were devised [3].
Despite numerous efforts, aqueous amine solutions still offer the
most economically efficient solution [59]. In turn, roomtemperature ionic liquids (RTILs), due to their tunability, emerge
as interesting competitors of aqueous amines [1,1012]. Mixtures
of amines and RTILs in water are currently under investigation,
since they allow to combine favorable physical chemical properties
of both [13,14].
RTILs are composed of bulky asymmetric carbon-based cations
and organic or inorganic anions. Most RTILs are non-volatile or
low-volatile [15,16], being thermally and electrochemically stable,
with high ignition points. These features minimize solvent losses
because of evaporation and decomposition at working conditions
Corresponding author.
E-mail address: vvchaban@gmail.com (V.V. Chaban).
http://dx.doi.org/10.1016/j.jct.2016.07.045
0021-9614/ 2016 Elsevier Ltd.

[4]. Non-volatile RTILs are also environmentally safe, since their


penetration into the living bodies is marginal. One must keep in
mind that viscosity decreases when RTILs are mixed with molecular co-solvents, such as absorbed CO2. Physical chemical properties
of RTILs viscosity, ionic and thermal conductivities, density, melting point, refraction coefficient, etc are readily modulated by
varying combinations of different cations and anions. RTILs exhibit
a number of possible physical and chemical interactions by a few
of their prospective coordination sites. Versatile interactions, such
as hydrogen bonding, dipole-dipole interaction, p-stacking of the
aromatic moieties are omnipresent in liquid RTILs [1723]. For
instance, the imidazolium-based RTILs exhibit moderate CO2
capture performance due to hydrogen bonding, which involves
the intrinsically acidic hydrogen atom at the C2 position of the
imidazole ring and the oxygen atom of CO2 [2426]. Pdua and
co-workers [27] experimentally proved that fluorination of the
imidazolium cations is significantly favorable for CO2 capture. The
experiments were conducted at low pressure and supplemented
by molecular simulations. It was found that the gas molecules prefer partially charged regions within the ionic solvent. In addition,
RTILs affinity to CO2 can be increased by grafting the amino group
ANH2 at a number of carbon atoms of the ring and the side chains

N.A. Andreeva, V.V. Chaban / J. Chem. Thermodynamics 103 (2016) 16

[2835]. Gupta reported insightful studies of CO2 capture by the


nitrile-based RTILs [11] and the tetracyanoborate-based RTILs [36].
Research efforts are underway to unite favorable properties of
RTILs and aqueous amines within the same substance. Chen and
coworkers [37] showed that the length of the aminoalkyl side
chain plays an important role to tune the CO2 performance of the
amino-functionalized imidazolium-based RTIL. Free energy of
adsorption, equilibrium constant, desorption temperature, reaction
speed, diffusion coefficients depend on the side chain. In this RTIL,
the cation works for gas capture, while the impact of the anion is
significantly weaker. The higher self-diffusion coefficients of CO2
were obtained in the cases of ethyl, propyl, and butyl side chains.
Lv and coworkers [38] demonstrated encouraging efficacy of the
amino-functionalized imidazolium-based RTIL. Similar conclusions
were earlier attained by Xue and coworkers [39]. Sairi and coworkers [40] investigated the pressure effect up to 30 bar proving that
CO2 capture increases significantly as higher pressures. This is in
line with theoretical expectations, according to the Le ChatelierBraun principle. As a result of chemisorption by the ANH2 group,
carbamates are formed, R2NACOOH, hence the number of molecules decreases twofold.
In the present work, we report ab initio calculations of the thermodynamic properties (enthalpy, entropy, Gibbs free energy),
vibrational frequencies, partial charges, and geometric parameters
to characterize CO2 capture by means of six amino-functionalized
organic cations following the carbamate formation pathway
(chemisorption). The organic cations were chosen based on a current interest to them in the field. While it is difficult to predict
which functionalized cation works best for CO2 capture, a systematical investigation of various RTIL families should be helpful.

2. Methodology
The wave functions were optimized at the Mller-Plesset second order level of theory [41]. The calculations were started from
conventional Hartree-Fock method, whose self-consistence field
convergence criterion was set to 10 8 Hartree. Subsequently, an
energy contribution originated from the correlation of electrons
was iteratively computed. Frequency analysis was conducted and
molecular partition functions were used to deduct thermodynamic
potentials for the products and the reactants [42]. The system
wave function was approximated by the atom-centered splitvalence triple-zeta polarized basis set with diffuse functions for
every atom, 6311++G. Prior to conducting analysis, each molecular configuration was optimized until the local minimum was
found. The iterations terminated, when the forces on all atoms,
including hydrogen atoms, became smaller than 0.4 kJ mol 1 nm 1. Only configurations with positive vibrational frequencies
were analyzed to exclude transition states.
The initial configurations of the simulated species were prepared in Gabedit [43]. VMD (Visual Molecular Dynamics, v. 1.9)
[44] was used to visualize results, measure geometric parameters,
and prepare artwork at high resolution. The GAMESS package was
applied using efficient parallelization over 16 cores [41].

Scheme 1. Cumulative reaction, which is responsible for CO2 fixation by the


aminated cation of RTIL. Carbamic acid is formed as a product.

3. Results and discussion


Chemisorption of CO2 takes place in accordance with Scheme 1.
The carbamate formation leads to decrease of total number of
molecules in the reaction mixture. In aqueous media, an acidic
environment should be expected, since the carboxyl group dissociates. Furthermore, the product is often depicted in the form of
zwitter-ion, which originates from transferring the hydrogen atom
of AC(O2)AH to the amino group NH+2. Such equilibrium is natural in polar solvents, but does not exist in the gas phase, as hereby
simulated. Furthermore, it is not important to rate different sites of
the cations, since the ionization contribution is equal in all cases.
The gas-phase thermodynamics does not account for important
physical effects, the most evident and impactful of which are solvation and interactions with the neighboring ions. Nevertheless, the
results are meaningful in most cases permitting to guide experiments. For instance, a relatively versatile test set involving 454
cases revealed that an average absolute deviation of the computational results from experiment is below 1 kcal mol 1 [45].
Fig. 1 depicts optimized (local-minimum) geometries of six simulated cations. These cations represent six different families of
RTILs. Each of the cations offers a number of sites (carbon atoms)
to be aminated. The aminated sites, subsequently, participate in
the CO2 chemisorption reaction following the same mechanism
as aqueous amines. The families of RTILs were selected based on
the pace of their investigation nowadays. With an exception of
tetraethylammonium, all cations contain a ring. The only cation
containing three different chemical elements in its ring is morpholinium. We also simulated thermodynamics of the carbamate
formation reaction for monoethanolamine (MEA) using the same
level of theory to get a reference point.
Standard thermodynamic potentials (Table 1) suggest that
chemisorption depends on the amination site within the cation.
Note that the absolute values of the reported potentials correspond
to simplified reactions (Scheme 1), which would occur in the gas
phase, rather than in an aqueous media. In particular, water fosters
formation of zwitter-ions, RANH+2ACOO , rather than protonated
carbamic acid, RANHACOOH. In both cases, R is an organic cation
of RTIL. Subsequently, ANACOO liberates CO2 in the form of HCO3
after its interaction with water (solvation). The detailed mechanisms are given in the classical works [4648]. The outlined difference should be kept in mind while comparing the reported
absolute values. The primary goal of the present work is to compare different organic cations and different amination sites within

Fig. 1. The optimized structures of the investigated organic cations: (1) tetraethylammonium (TEA+, (C2H5)4N+); (2) 1-ethyl-3-methylimidazolium (EMIM+,
C3H3N2(CH3)(C2H5)+);
(3)
1-methylmorpholinium
(MM+,
C4H9NO(CH3)+);
(4) 1-methylpyridinium (MPD+, C5H5N(CH3)+); (5) 1-ethyl-1-methylpyrrolidinium
(EMPL+,
C4H8N(CH3)(C2H5)+);
(6)
1-ethyl-2-methylpyrazolium
(EMPZ+,
C3H3N2(CH3)(C2H5)+). The labeled sites were used to attach the amino group,
which subsequently gave rise to a respective carbamate molecule.

N.A. Andreeva, V.V. Chaban / J. Chem. Thermodynamics 103 (2016) 16


Table 1
Standard thermodynamic potentials computed from molecular partition functions for
CO2 chemisorption via the carbamate formation of the six organic cations of RTILs. As
a reference reaction, we used chemisorption by MEA.
Atom
TEA
ET

DH, kJ mol

TDS, kJ mol

DG, kJ mol

13.2

46.7

59.9

EMIM+
C3
CW1
CW2
CH2
ET
ME

41.0
39.7
86.3
46.1
15.9
11.3

41.6
45.2
46.4
45.7
44.7
45.0

82.6
84.6
132.7
91.7
60.6
56.3

MM+
C1
C2
C3
C4
ME

23.0
85.0
42.7
30.8
71.5

47.2
44.1
44.2
45.2
46.8

70.2
129.1
86.9
76.0
118.3

MPD+
C1
C2
C3
ME

65.6
105.8
61.6
37.9

46.6
43.6
43.6
52.0

112.1
149.3
105.2
90.0

EMPL+
C1
C2
C3
C4
CH2
ET
ME

103.2
36.2
37.7
36.0
51.2
1.6
30.3

47.0
46.9
45.5
45.7
47.5
47.6
47.4

150.2
83.0
83.3
81.7
98.7
49.3
77.8

EMPZ+
C1
C2
C3
CH2
ET
ME

64.3
32.1
45.3
93.9
8.1
34.6

49.7
36.3
45.5
45.4
44.0
44.1

114.0
68.4
90.8
139.3
52.0
78.7

MEA
C

39.8

43.7

83.5

these organic cations, based on their performance for CO2 capture.


This goal does not require reproduction of the experimental thermodynamic results, but strongly requires accurate simulation of
the electronic density within each cation. Therefore, an electroncorrelation ab initio method, MP2, was used to obtain wave functions for frequency analysis.
The Gibbs free energy characterizing chemisorption of CO2 by
MEA amounts to 83.5 kJ mol 1. Interestingly, this is not the most

favorable reaction as one would expect. The amino group linked


at any ET sites of the organic cation offers 5060 kJ mol 1. This
observation is not influenced by aromaticity of the ring. In a few
cases, the ME site is also more favorable than the reference system
of MEA. In turn, the sites of the rings are generally less favorable.
Such ordering of sites inspires synthetic efforts to obtain specific
derivatives of RTILs for CO2 capture.
The thermodynamic potentials can be rationalized in the framework of the well-established concept describing reactions between
nucleophile and electrophile. CO2 is an electrophile with electron
deficiency on carbon. In turn, ANH2 is a nucleophile. The nucleophilicity of ANH2 can be increased by the neighboring electron
donating groups. The amino group of MEA neighbors the hydroxyl
group, which shifts electron density towards itself. In turn, the
amino group at the ET site does not have such a group and retains
its own electron density. The thermodynamic computations capture this difference accurately suggesting more favorable CO2
attachment to ET. Based on the same considerations, the ME sites
are quite favorable, but less favorable than ET in the investigated
cations. The sites within the ring are generally less favorable,
since the rings of the organic cations are electron deficient. The
nucleophilicity can be quantified by means of partial electron
charges computed according to the Hirshfeld definition [49],
the Hipart program (molmod.ugent.be/software). Compare,
q(ET) = (0.23  0.24)e to q(ME) = (0.21  0.19)e in different
organic cations. The nitrogen atoms linked to the carbon atoms
of the ring can also be electron rich. For instance, the C3 site of
EMPL+ carries 0.23e, resulting is a relatively favorable chemisorption reaction, 83.3 kJ mol 1. The nucleophilicity of the amino
groups linked to the sites within the ring generally depends on
their location with respect to heteroatoms (oxygen, nitrogen) and
on whether the ring is aromatic.
All carbamate formation reactions depend greatly on the entropic factor (Fig. 2), whose contribution exceeds 90% in some cases.
An important role of entropy is natural in the reactions, in which

Table 2
Angle CANAC (see Fig. 2) for different products of the CO2 chemisorption. The
analogous angle in MEA equals to 123.8.
Atom

TEA+

EMIM+

MM+

MPD+

EMPL+

EMPZ+

C1/CW1
C2/CW2
C3
C4
CH2
ET
ME

123.8

122.9
121.9
123.3

124.7
123.2
118.2

122.4
124.0
121.7
122.8

118.3

122.9
126.0
132.1

122.0

125.6
127.2
127.4
125.2
128.9
124.2
124.9

116.7
128.1
122.6

115.2
123.5
122.6

Fig. 2. Fraction of the entropic contribution to the Gibbs free energy of the CO2 chemisorption.

N.A. Andreeva, V.V. Chaban / J. Chem. Thermodynamics 103 (2016) 16

gases are formed out of the non-gaseous reactants or vice versa.


The yield of such reactions can be modulated by regulating pressure. Indeed, it is experimentally known that CO2 capture strongly
depends on the CO2 partial pressure [50].
The carbamate moiety occupies a significant region of space;
therefore steric considerations may be important in some cases.
Table 2 summarizes angles, which were computed using the car-

Fig. 3. Optimized structures of carbamates obtained from (a) MEA and (b) EMPZ+.
The CANAC angle characterizes the configuration of the grafted carbamate moiety.

bon and nitrogen atoms belonging to the carbamate group and


the carbon atom belonging to the radical (organic cation), as exemplified in Fig. 3. We used an analogous angle in MEA as a reference
assuming that no steric hindrances exist in that case. In turn, the
mutual location of the carbamate moiety and the hydrocarbon side
chains need to be examined.
In the cases of the ET sites, the CANAC angles are nearly the
same as in MEA (Table 2). The largest perturbations of the angle
are observed at the CH2 site. It is particularly true when two side
chains are connected to the same atom (nitrogen) of the ring
(e.g. EMPL+) or to the neighboring nitrogen atoms (e.g. EMPZ+). In
turn, the CH2 site in EMIM+ exhibits quite a natural value, 124.7.
Another interesting case is the C1 site in EMPZ+, where the angle
perturbation correlates with the thermodynamic potentials. Analysis of the CANAC angle helps explain the cases, where the electron density distribution considerations are not sufficient.
Attachment of the carbamate moiety to the organic cations
results in a very intensive vibrational peak at ca. 1900 cm 1
(Fig. 4). This vibration corresponds to antisymmetric stretching in
the C@O group. High intensity is due to polarity of this group
and, hence, the largest changes of the dipole moment upon vibration. The differences in the location of this peak are marginally

Fig. 4. Mid- and far-infrared vibrational spectra of the EMPL+ cation based carbamates. See Fig. 1 for site designation.

N.A. Andreeva, V.V. Chaban / J. Chem. Thermodynamics 103 (2016) 16


Table 3
Computed frequencies and intensities for the antisymmetric vibrations of C@O in
different cations.
Site

Frequency,
cm 1

Intensity,
km mol 1

Site

Frequency,
cm 1

Intensity,
km mol 1

EMIM+
C3
CW1
CW2
CH2
ET
ME

1880
1878
1881
1865
1869
1817

605
648
450
603
711
369

MM+
C1
C2
C3
C4
ME

1892
1881
1872
1891
1854

622
541
602
632
273

EMPL+
C1
C2
C3
C4
CH2
ET
ME

1855
1868
1868
1870
1861
1862
1869

330
707
714
638
605
643
570

EMPZ+
C1
1865
C2
1872
C3
1874
CH2 1850
ET
1874
ME
1878

259
795
642
276
719
609

MPD+
C1
C2
C3
ME

1879
1885
1872
1883

646
495
697
620

TEA+
ET
1867
MEA
C
1850

714
556

dependent on the identity of the organic cation and the aminofunctionalized site within the cation (Table 3). The frequencies at
the ET sites are higher, as compared to those in MEA. The computed spectra suggest that spectroscopic identification of the CO2
capture is straightforward.

4. Conclusions
To recapitulate, the present paper reports ab initio thermodynamic analysis of the performance of different RTILs upon
chemisorption of CO2 via the carbamate formation mechanism.
The regularities of chemisorption were corroborated by means of
computational infrared spectroscopy, geometric parameters, and
charge density distributions. The CO2 capture was found to significantly depend on the location of the amino group within the
organic cation. In terms of Gibbs free energy, the difference may
exceed 100%. Interestingly, the performance of cations with the
amino-functionalized side chains is better than that of MEA. In
all cases, the entropic contribution constitutes a very important
fraction of the Gibbs free energy. Thus, the investigated reactions
are sensitive to pressure.
Thermodynamic potentials correlate with the electron density
distribution and can be explained as reactions between the electrophile (CO2) and a set of nucleophiles (amino-functionalized
organic cations) of somewhat different strength. This strength
depends on the carbon atom of the cation, to which the amino
group is grafted. A less important factor is steric hindrances due
to formation of the carbamate moiety in the vicinity of the side
chains.
A surprising result is that hydrocarbon chains were found to
exhibit smaller Gibbs free energies, as compared to MEA. While
the amino-functionalized hydrocarbons R-(NH2)n cannot be
directly used for CO2 capture due to their notorious volatility, they
lose volatility completely when grafted to the organic cations. Furthermore, many RTILs are soluble in water, unlike hydrocarbons.
Usage of the aminated RTILs, therefore, allows for avoiding hydroxyl groups, whose goal in the industrial amines (MEA, methyl
diethanolamine, etc) is to modulate solubility and decrease volatility. The computational results reported in the present work inspire
development of new CO2 scavengers, whose working characteris-

tics may be comparable (or even better, e.g. because of nonvolatility) than those of aqueous amines.
References
[1] S. Babamohammadi, A. Shamiri, M.K. Aroua, A review of CO2 capture by
absorption in ionic liquid-based solvents, Rev. Chem. Eng. 31 (2015) 383412.
[2] C. Chen, J. Kim, W.S. Ahn, CO2 capture by amine-functionalized nanoporous
materials: a review, Korean J. Chem. Eng. 31 (2014) 19191934.
[3] C.H. Yu, C.H. Huang, C.S. Tan, A review of CO2 capture by absorption and
adsorption, Aerosol Air Qual. Res. 12 (2012) 745769.
[4] C. Gouedard, D. Picq, F. Launay, P.L. Carrette, Amine degradation in CO2
capture. I. A review, Int. J. Greenhouse Gases Control 10 (2012) 244270.
[5] M. Sobrino, E.I. Concepcin, . Gmez-Hernndez, M.C. Martn, J.J. Segovia,
Viscosity and density measurements of aqueous amines at high pressures:
MDEA-water and MEA-water mixtures for CO2 capture, J. Chem. Thermodyn.
98 (2016) 231241.
[6] H. Pahlavanzadeh, S. Fakouri Baygi, Modeling CO2 solubility in aqueous
methyldiethanolamine solutions by perturbed Chain-SAFT Equation of State, J.
Chem. Thermodyn. 59 (2013) 214221.
[7] I.M.S. Lampreia, .F.S. Santos, M.-L.C.J. Moita, L.C.S. Nobre, Thermodynamic
study of aqueous 2-(isopropylamino)ethanol. A sterically hindered new amine
absorbent for CO2 capture, J. Chem. Thermodyn. 81 (2015) 167176.
[8] D. Fu, L. Wang, P. Zhang, C. Mi, Solubility and viscosity for CO2 capture process
using MEA promoted DEAE aqueous solution, J. Chem. Thermodyn. 95 (2016)
136141.
[9] D. Fernandes, W. Conway, R. Burns, G. Lawrance, M. Maeder, G. Puxty,
Investigations of primary and secondary amine carbamate stability by 1H NMR
spectroscopy for post combustion capture of carbon dioxide, J. Chem.
Thermodyn. 54 (2012) 183191.
[10] O. Ben Ghanem, M.I.A. Mutalib, J.M. Leveque, G. Gonfa, C.F. Kait, M. El-Harbawi,
Studies on the physicochemical properties of ionic liquids based On 1-Octyl-3methylimidazolium Amino Acids, J. Chem. Eng. Data 60 (2015) 17561763.
[11] K.M. Gupta, J.W. Jiang, Systematic investigation of nitrile based ionic liquids
for CO2 capture: a combination of molecular simulation and ab initio
calculation, J. Phys. Chem. C 118 (2014) 31103118.
[12] M. Klahn, A. Seduraman, What determines CO2 solubility in ionic liquids? A
molecular simulation study, J. Phys. Chem. B 119 (2015) 1006610078.
[13] P. Zhang, L.X. Du, D. Fu, Experiment and model for the surface tension of
amine-ionic liquids aqueous solutions, J. Chem. Thermodyn. 74 (2014) 16.
[14] F.L. Bernard, F.D. Vecchia, M.F. Rojas, R. Ligabue, M.O. Vieira, E.M. Costa, V.V.
Chaban, S. Einloft, Anticorrosion protection by amine-ionic liquid mixtures:
experiments and simulations, J. Chem. Eng. Data 61 (2016) 18031810.
[15] J.M.S.S. Esperanca, J.N.C. Lopes, M. Tariq, L.M.N.B.F. Santos, J.W. Magee, L.P.N.
Rebelo, Volatility of aprotic ionic liquids a review, J. Chem. Eng. Data 55
(2010) 312.
[16] L.P.N. Rebelo, J.N.C. Lopes, J.M.S.S. Esperanca, E. Filipe, On the critical
temperature, normal boiling point, and vapor pressure of ionic liquids, J.
Phys. Chem. B 109 (2005) 60406043.
[17] D. Roy, M. Maroncelli, Simulations of solvation and solvation dynamics in an
idealized ionic liquid model, J. Phys. Chem. B 116 (2012) 59515970.
[18] K. Paduszynski, E.V. Lukoshko, M. Krolikowski, U. Domanska, J. Szydlowski,
Thermodynamic study of binary mixtures of 1-butyl-1-methylpyrrolidinium
dicyanamide ionic liquid with molecular solvents: new experimental data and
modeling with PC-SAFT equation of state, J. Phys. Chem. B 119 (2015) 543
551.
[19] X.X. Zhang, M. Liang, N.P. Ernsting, M. Maroncelli, Conductivity and solvation
dynamics in ionic liquids, J. Phys. Chem. Lett. 4 (2013) 12051210.
[20] K. Paduszynski, M. Okuniewski, U. Domanska, Renewable feedstocks in green
solvents: thermodynamic study on phase diagrams of D-Sorbitol and xylitol
with dicyanamide based ionic liquids, J. Phys. Chem. B 117 (2013) 70347046.
[21] X.X. Zhang, J. Breffke, N.P. Ernsting, M. Maroncelli, Observations of probe
dependence of the solvation dynamics in ionic liquids, Phys. Chem. Chem.
Phys. 17 (2015) 1294912956.
[22] C.E.S. Bernardes, T. Mochida, J.N.C. Lopes, Modeling the structure and
thermodynamics of ferrocenium-based ionic liquids, Phys. Chem. Chem.
Phys. 17 (2015) 1020010208.
[23] K. Paduszynski, U. Domanska, Solubility of aliphatic hydrocarbons in
piperidinium ionic liquids: measurements and modeling in terms of
perturbed-chain statistical associating fluid theory and nonrandom
hydrogen-bonding theory, J. Phys. Chem. B 115 (2011) 1253712548.
[24] V.V. Chaban, O.V. Prezhdo, Ionic and molecular liquids: working together for
robust engineering, J. Phys. Chem. Lett. 4 (2013) 14231431.
[25] D. Almantariotis, S. Stevanovic, O. Fandino, A.S. Pensado, A.A.H. Padua, J.Y.
Coxam, M.F.C. Gomes, Absorption of carbon dioxide, nitrous oxide, ethane and
nitrogen by 1-Alkyl-3-methylimidazolium (C(n)mim, n = 2,4,6) Tris
(pentafluoroethyl)trifluorophosphate Ionic Liquids (eFAP), J. Phys. Chem. B
116 (2012) 77287738.
[26] S. Stevanovic, A. Podgorsek, A.A.H. Padua, M.F.C. Gomes, Effect of water on the
carbon dioxide absorption by 1-Alkyl-3-methylimidazolium acetate ionic
liquids, J. Phys. Chem. B 116 (2012) 1441614425.
[27] D. Almantariotis, T. Gefflaut, A.A.H. Padua, J.Y. Coxam, M.F.C. Gomes, Effect of
fluorination and size of the alkyl side-chain on the solubility of carbon dioxide
in 1-Alkyl-3-methylimidazolium Bis(trifluoromethylsulfonyl)amide Ionic
Liquids, J. Phys. Chem. B 114 (2010) 36083617.

N.A. Andreeva, V.V. Chaban / J. Chem. Thermodynamics 103 (2016) 16

[28] X.Y. Luo, X. Fan, G.L. Shi, H.R. Li, C.M. Wang, Decreasing the viscosity in CO2
capture by amino-functionalized ionic liquids through the formation of
intramolecular hydrogen bond, J. Phys. Chem. B 120 (2016) 28072813.
[29] B.H. Lv, Y.F. Xia, Y. Shi, N. Liu, W. Li, S.J. Li, A novel hydrophilic amino acid ionic
liquid [C(2)OHmim][Gly] as aqueous sorbent for CO2 capture, Int. J.
Greenhouse Gas Control 46 (2016) 16.
[30] J. Li, Z.D. Dai, M. Usman, Z.W. Qi, L.Y. Deng, CO2/H-2 separation by amino-acid
ionic liquids with polyethylene glycol as co-solvent, Int. J. Greenhouse Gas
Control 45 (2016) 207215.
[31] V. Hiremath, A.H. Jadhav, H. Lee, S. Kwon, J.G. Seo, Highly reversible CO2
capture using amino acid functionalized ionic liquids immobilized on
mesoporous silica, Chem. Eng. J. 287 (2016) 602617.
[32] B.B. Cao, J.Y. Du, S.Y. Liu, X. Zhu, X.J. Sun, H.T. Sun, H. Fu, Carbon dioxide
capture by amino-functionalized ionic liquids: DFT based theoretical analysis
substantiated by FT-IR investigation, RSC Adv. 6 (2016) 1046210470.
[33] Y.S. Sistla, A. Khanna, CO2 absorption studies in amino acid-anion based ionic
liquids, Chem. Eng. J. 273 (2015) 268276.
[34] Z.M. Xue, Y.W. Zhang, X.Q. Zhou, Y.Y. Cao, T.C. Mu, Thermal stabilities and
decomposition mechanism of amino- and hydroxyl-functionalized ionic
liquids, Thermochim. Acta 578 (2014) 5967.
[35] Y. Zhang, P. Yu, Y.N. Luo, Absorption of CO2 by amino acid-functionalized and
traditional dicationic ionic liquids: properties, Henrys law constants and
mechanisms, Chem. Eng. J. 214 (2013) 355363.
[36] K.M. Gupta, Tetracyanoborate based ionic liquids for CO2 capture: from
ab initio calculations to molecular simulations, Fluid Phase Equilib. 415 (2016)
3441.
[37] J.J. Chen, W.W. Li, X.L. Li, H.Q. Yu, Carbon dioxide capture by aminoalkyl
imidazolium-based ionic liquid: a computational investigation, Phys. Chem.
Chem. Phys. 14 (2012) 45894596.
[38] B.H. Lv, G.H. Jing, Y.H. Qian, Z.M. Zhou, An efficient absorbent of amine-based
amino acid-functionalized ionic liquids for CO2 capture: high capacity and
regeneration ability, Chem. Eng. J. 289 (2016) 212218.

[39] Z.M. Xue, Z.F. Zhang, J. Han, Y. Chen, T.C. Mu, Carbon dioxide capture by a dual
amino ionic liquid with amino-functionalized imidazolium cation and taurine
anion, Int. J. Greenhouse Gas Control 5 (2011) 628633.
[40] N.A. Sairi, R. Yusoff, Y. Alias, M.K. Aroua, Solubilities of CO2 in aqueous Nmethyldiethanolamine and guanidinium trifluoromethanesulfonate ionic
liquid systems at elevated pressures, Fluid Phase Equilib. 300 (2011) 8994.
[41] M.W. Schmidt, K.K. Baldridge, J.A. Boatz, S.T. Elbert, M.S. Gordon, J.H. Jensen, S.
Koseki, N. Matsunaga, K.A. Nguyen, S. Su, et al., General atomic and molecular
electronic structure system, J. Comput. Chem. 14 (1993) 13471363.
[42] H.A. Guggeheim, Thermodynamics: An Advanced Treatment for Chemists and
Physicists, fifth ed., Elsevier, New York, 1967.
[43] A.R. Allouche, Gabedit-A graphical user interface for computational chemistry
softwares, J. Comput. Chem. 32 (2011) 174182.
[44] W. Humphrey, A. Dalke, K.V.M.D. Schulten, Visual molecular dynamics, J. Mol.
Graphics 14 (1996) 3338.
[45] L.A. Curtiss, P.C. Redfern, K. Raghavachari, Gaussian-4 theory, J. Chem. Phys.
126 (2007) 084108.
[46] S. Bishnoi, G.T. Rochelle, Absorption of carbon dioxide in aqueous piperazine/
methyldiethanolamine, AIChE J. 48 (2002) 27882799.
[47] H. Hikita, S. Asai, Y. Katsu, S. Ikuno, Absorption of carbon-dioxide into aqueous
monoethanolamine solutions, AIChE J. 25 (1979) 793800.
[48] E. Sada, H. Kumazawa, M.A. Butt, Chemical absorption kinetics over a widerange of contact time absorption of carbon-dioxide into aqueous-solutions of
monoethanolamine, AIChE J. 22 (1976) 196198.
[49] F.L. Hirshfeld, Bonded-atom fragments for describing molecular charge
densities, Theoret. Chim. Acta 44 (1977) 129138.
[50] A.S. Aquino, F.L. Bernard, J.V. Borges, L. Mafra, F. Dalla Vecchia, M.O. Vieira, R.
Ligabue, M. Seferin, V.V. Chaban, E.J. Cabrita, et al., Rationalizing the role of the
anion in CO2 capture and conversion using imidazolium-based ionic liquid
modified mesoporous silica, RSC Adv. 5 (2015) 6422064227.

JCT 16-449

Vous aimerez peut-être aussi