Sie sind auf Seite 1von 11

Journal of CO₂ Utilization 34 (2019) 282–292

Contents lists available at ScienceDirect

Journal of CO2 Utilization


journal homepage: www.elsevier.com/locate/jcou

Pyridine-containing ionic liquids lowly loaded in large mesoporous silica T


and their rapid CO2 gas adsorption at low partial pressure

Chunfeng Xuea, , Lin Fenga, Hongye Zhua, Ruichao Huanga, Enyang Wanga, Xiao Dua, Gang Liua,

Xiaogang Haoa, , Kaixi Lib
a
Department of Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, Shanxi, PR China
b
Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, Shanxi, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: To achieve low cost, high rate and attractive capacity of CO2 adsorption by using ionic liquid (IL), tetra-
Ionic liquid butylammonium 2-hydroxypyridine ([N4444][2-Op]) and tetrabutylphosphonium 2-hydroxypyridine ([P4444]
Mesopore [2-Op]) are newly prepared and immobilized into mesoporous silica (MS) with pore diameter of about 13.5 nm
CO2 capture for fabricating mesostructured ionogel. It takes only 10.0 min for the ionogel NMS-15 (MS loaded with 15% IL
Multiple active site
[N4444][2-Op]) to achieve an adsorption capacity of 1.081 mmol CO2/(g ionogel) (that is 1.10 mol CO2/(mol IL))
Silica
at 50.0 °C and low CO2 partial pressure. Also, it only takes 7.5 min for the ionogel PMS-10 (MS loaded with 10%
IL [P4444][2-Op]) to achieve 90% saturated adsorption capacity of 1.678 mmol CO2/(g ionogel) (that is 3.72 mol
CO2/(mol IL)), which is 12.76 times higher than that of the pure IL [P4444][2-Op]. Its adsorption capacity
retention is around 90% after performing 10 cycles continuous test. The results can be ascribed to the meso-
porous silica with large pore that benefits to not only load more exposed IL but also remain necessary pathway
for CO2 diffusion. Combining with low loading and cost, swift adsorption, high capacity as well as good cyclic
and thermal stability make the ionogel PMS-10 a competitive candidate in CO2 capture from the flue gas.

1. Introduction efforts had been paid to enhance their adsorption capacity and rate.
One feasible strategy was to introduce more active site into the ILs. For
To capture CO2 from the combustion of fossil fuel were funda- example, dual or multi active sites were introduced to acquire pro-
mentally important for alleviating the air pollution and global warming mising CO2 adsorption capacity [14,17]. By dispersing ILs into porous
[1–5]. On 8th October 2018, Intergovernmental Panel on Climate materials, synergistic effect had also been observed on the improve-
Change (IPCC) released an IPCC special report that also mentioned the ment of CO2 adsorption efficiency basing on the enlarged gas-liquid
strengthening the global response to the threat of climate change. interface [18–24]. In view of this, different porous materials had been
Current capture technologies based on organic amines were featured used to load ILs. Mesoporous polymethylmethacrylate microspheres
with energy intensive and accompanied with obvious loss of volatile impregnated with amino acid ILs showed CO2 adsorption capacity
components [6–8]. In view of this, solid materials including carbon [3], 1.30–1.53 mmol/g at 40.0 °C [25,26]. High adsorption capacities were
polymer [9], and metal organic framework were regarded as promising also observed on mesoporous silica supported hydroxyalkyl amidines
alternatives [3,10–12]. They showed acceptable CO2 adsorption capa- [27]. The silica surface-confined zinc-functionalized ILs took about 6 h
cities especially at given temperature and high pressure. But they were to achieve adsorption capacity of 2.0 mmol/g at 40.0 °C [28]. However,
difficult to be used in the practical industries. In recent years, various the tetrabutylphosphonium amino acid supported on silica gel only
ionic liquid (IL) had been designed for capturing CO2 due to their showed 50% adsorption capacity of the pure IL [29]. The adsorption
negligible volatility, and tunable properties [8,13–15]. It took a long processes mentioned above were as slow as the pure ILs. Similar result
time (several hours) for them to achieve the adsorption balance, which was observed in mesoporous silica MCM-41 loaded with IL [P66614][2-
mainly results from the high viscosity, low diffusion coefficient of ILs Op] [30]. “Inverse” supported IL phase showed enhanced adsorption
[8,16,17]. kinetics but a poor thermal stability [31]. Obviously, mass transport
To economically make full use of the ILs for CO2 capture, many limitation in most of the above ionogels will inhibit their practical


Corresponding authors.
E-mail addresses: cfxue@fudan.edu.cn (C. Xue), xghao@tyut.edu.cn (X. Hao).

https://doi.org/10.1016/j.jcou.2019.06.015
Received 23 March 2019; Received in revised form 6 June 2019; Accepted 17 June 2019
2212-9820/ © 2019 Elsevier Ltd. All rights reserved.
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

applications. Comparing with the supported ILs with one active site, ILs
with dual active sites may achieve higher adsorption capacity in a short
time basing on allosteric effect [20,32,33]. We expect that the sup-
porter with large pore size is a good platform for loading the IL in a
monomolecular layer and necessary diffusion pathway for CO2 to
achieve higher and faster adsorption. Thus, it is desirable that an io-
nogel can be reasonably designed by loading multi-active-site con-
taining IL into a supporter with large pores.
Herein, mesoporous silica with pore diameter of 13.5 nm is pre-
pared as the support, which can provide huge nanospace for loading IL
and interconnected highway for CO2 diffusion. Two kinds of 2-hydro-
xypyridine anion functionalized ILs are designed after considering the
multiple-site cooperative interactions of pyridine-containing anion and
the small free volume of cation [P4444]+ and [N4444]+. The obtained
ionogels PMS-w and NMS-w exhibit high adsorption capacities in short
time. With 10% IL [P4444][2-Op] filling in ionogel PMS-10, the highest
adsorption capacity of 1.678 mmol CO2/(g ionogel) is observed at
50.0 °C in a binary mixture. It takes no more than 10.0 min for the io-
nogel PMS-10 to achieve 90% of saturated adsorption capacity due to
Fig. 1. 1H NMR of IL [N4444][2-Op].
the fast gas sorption on the electronegative oxygen and nitrogen atoms
of the IL, which is fairly faster than the bulk ILs and reported ones. The
adsorption capacity retention of about 90% is observed after per- Compared with the viscosity of water (0.55 mPa s), its viscosity at
forming 10 cycles of adsorption-desorption tests. Also, an enhanced 20.0 °C was measured to be 982.0 mPa s, showing a fairly high viscosity.
thermal stability up to 430.0 °C in N2 is observed for the low loading Its molecular weight was calculated at 336.53. Its water content is
ionogels. The ionogel NMS-15 shows the highest adsorption capacity of measured at 1.24%. 1H NMR (500.0 MHz, DMSO, 25.0 °C) of the IL
1.081 mmol CO2/(g ionogel) among all the ionogels NMS-w in [N4444][2-Op]: δ = 0.93, 1.25–1.37, 1.50–1.63, 5.75, 6.89, and
11.0 min. 7.56 ppm (Fig. 1). 13C NMR of IL [N4444][2-Op]: δ = 13.97, 16.69,
23.54, 39.98, 52.04, 58.04, 103.74, 114.50, 129.25, 133.39, 136.60,
2. Experimental 136.23, 147.70, 172.36 ppm (Fig. 2(a)). Five peaks at 103.74, 114.50,
136.23, 147.70, and 172.36 ppm are replaced with the ones at 105.03,
2.1. Chemicals 117.47, 138.99, 141.67, and 167.17 ppm after the CO2 adsorption
(Fig. 2(b)). CO2 adsorption capacity of the bulk IL [N4444][2-Op] was
Tetrabutylammonium hydroxide solution ([N4444][OH], 40% in measured at 1.541 mmol CO2/(g IL) by bubbling pure CO2 at 50.0 °C for
H2O), tetrabutylphosphonium hydroxide solution ([P4444][OH], 40% in more than 60.0 min (Fig. 3).
H2O) and 2-hydroxypyridine (2-Op) were purchased on J&K Scientific The IL [P4444][2-Op] was prepared following the same process.
Ltd. Fumed silica (SiO2), Trimethylbenzene (TMB), cetyltrimethyl am- Briefly, 9.7744 g [P4444][OH] was dissolved in 20.0 mL EtOH and
monium bromide (CTAB), sodium hydroxide (NaOH), and anhydrous 1.3450 g 2-Op was dissolved in another 20.0 mL EtOH. Then the two
ethanol (EtOH) were purchased from Tianjin Zhiyuan Chemical solutions were mixed at 30.0 °C for 24.0 h and evaporated at 50.0 °C for
Reagent Co., Ltd. N2 (> 99.99%), and CO2 (> 99.99%) were bought 1.0 h. After further dried at 60.0 °C under vacuum for 2.0 d, the IL
from Taiyuan Taineng Gas Co. Ltd. [P4444][2-Op] is formed finally. Its viscosity was measured to be
202.05 mPa s at 20.0 °C. Its molecular weight is calculated at 353.53.
2.2. Preparation of mesoporous silica (MS) Besides, 13C NMR (500 MHz, DMSO, 25.0 °C) of IL [P4444][2-Op]:
δ = 13.3, 14.1, 18.6, 18.9, 23.6, 23.7, 23.8, 23.9, 24.3, 27.3, 27.8,
The molar ratio of raw materials for synthesizing MS was selected as 105.5, 120.1, 132.2, 139.7, 160.4 ppm (Fig. 4(a)). New peaks 164.2 and
followings: SiO2: 1.3 TMB: 0.25 NaOH: 0.10 CTAB: 20.0 H2O [34].
Typically, 1.2270 g CTAB was firstly dissolved and stirred in 12.0 g
H2O. Then 0.3340 g NaOH was added with further stirring at room
temperature to get a clear solution. 5.2078 g TMB was also added. Fi-
nally, 2.0 g fumed SiO2 was added with strong stirring for 30.0 min. The
obtained mixture was transferred into autoclave and heated at 70.0 °C
for 8.0 h. The white solid was washed with distilled water for three
times after the autoclave was cooled to room temperature. The washed
sample was dried at 115.0 °C for 12.0 h. Then the sample was calcined
at 500.0 °C for 8.0 h in muffle oven to remove the remained CTAB and
TMB. The ramp rate of 2.0 °C/min is used for the calcination. The ob-
tained sample was denoted as MS.

2.3. Preparation of pyridine-containing anion functionalized ILs

A new IL [N4444][2-Op] was prepared as followings. 9.6376 g


[N4444][OH] was dissolved in 20.0 mL EtOH and stirred at room tem-
perature. Then, 1.4130 g 2-Op was dissolved in another 20.0 mL EtOH.
Subsequently, the two solutions were mixed together at 30.0 °C for
24.0 h and evaporated in rotary evaporator at 50.0 °C for 1.0 h to re-
move EtOH. After further dried at 60.0 °C under vacuum for 48.0 h to
13
remove the trace H2O, the IL [N4444][2-Op] was prepared finally. Fig. 2. C NMR of IL [N4444][2-Op] before (a) and after (b) CO2 adsorption.

283
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

135.6 ppm ascribed to O-CO2 and N-CO2 appear after the CO2 adsorp-
tion (Fig. 4(b)). 31P NMR of the IL: δ = 33.8 ppm. 1H NMR of the IL:
δ = 0.90, 1.36–1.51, 2.16–2.25, 5.71, 6.86, and 7.56 ppm. Its CO2 ad-
sorption capacity of 1.61 mmol CO2/(g IL) was measured by bubbling
CO2 into it at 50.0 °C for more than 100.0 min (Fig. 5).

2.4. Preparation of ionogels

The sample MS was used as the supporter for loading [N4444][2-Op]


and [P4444][2-Op]. In a typical run, a given amount of the dried IL was
dissolved in 30.0 g of EtOH and stirred at room temperature for 1.0 h.
Then, 1.0 g MS was added into the solution and heated at 80.0 °C for
2.0 h. Finally, the mixture is further dried at 100.0 °C under vacuum
overnight to remove the trace H2O and EtOH. According to the loading
amount of the IL, obtained ionogels [N4444][2-Op]@MS or [P4444][2-
Op]@MS were denoted as NMS-w and PMS-w, where “w” means weight
percentage of the IL in the initial mixture. Here, we prepared the io-
nogels NMS-5, NMS-10, NMS-15, NMS-30, PMS-1, PMS-3, PMS-5, PMS-
Fig. 3. CO2 adsorption curve of bulk IL [N4444][2-Op] bubbled with pure CO2 at 10, PMS-15, and PMS-30 respectively.
50.0 °C.
2.5. Characterization

N2 adsorption-desorption isotherms were measured on JW-BK122W


specific surface area and pore size analyzer (Beijing JWGB Sci. & Tech.
Co., Ltd.) at −196.0 °C. The sample MS (100.0–200.0 mg) was degassed
at 300.0 °C for 3.0 h under vacuum before the measurement, while io-
nogels (100.0–200.0 mg) were degassed at 120.0 °C for 8.0 h. Specific
surface area of samples were calculated in the range of P/P0 = 0.05–0.2
by using Brunauer–Emmett–Teller (BET) method. Pore size distribu-
tions were calculated with Barrett–Joyner–Halenda (BJH) method
based on adsorption branches of the isotherms. X-ray diffraction (XRD)
patterns of samples were collected on a Rigaku Ultima IV type X-ray
diffractometer, equipped with the Cu Kα radiation, (λ = 0.154056 nm,
40.0 kV, 40.0 mA) with a scan range of 1.0–7.0°. Fourier transfer in-
frared spectroscopy (FTIR) spectrum of all the samples were recorded
with a solution of 2.0 cm–1 in the range of 400.0–4000.0 cm–1 on a
Thermo Nicolet-360 FTIR spectrometer. NMR spectra of ILs were re-
corded on Bruker DRX500.0 MHz. The viscosity was measured on
Brookfield DV2T viscometer with a 53 mm spindle. The H2O content of
the sample was measured on AKF-3 Karl Fischer coulometer.
Thermogravimetry (TG) curves of samples were collected on a Netzsch
13
Fig. 4. C NMR of IL [P4444][2-Op] before (a) and after (b) the CO2 adsorption. STA 449 F5 TG/DTG instrument. About 7.0 mg of samples were placed
onto an alumina crucible and heated at a ramp of 10.0 °C/min in the
range of from 30.0 to 800.0 °C in N2 flow (100.0 mL/min).

2.6. CO2 adsorption measurement

CO2 adsorption capacities of ionogels were calculated by recording


breakthrough curves on Gasboard-3100 gas analyzer as shown in Fig. 6.
The CO2 and N2 were monitored by mass flow controller DO7-11C se-
parately. 1.0 g dried ionogel was placed into the sample tube with an
inner diameter of 10.0 mm and a length of 220.0 mm. First, the sample
was activated at 120.0 °C for 2.0 h in N2 flow of 98.5 mL/min. Then it
was cooled to given adsorption temperature of 50.0 °C, which referred
to typical temperature of the flue gas. Next, a mixed gas CO2/N2 was
introduced via bypass until the concentration of CO2 was constant, in
which the volume concentration of CO2 was approximate 14.9 vol.% in
N2 balancing at a flow rate of 98.5 mL/min. Afterwards, the mixed gas
CO2/N2 was switched to the sample tube at 50.0 °C to perform the
adsorption experiment. When the CO2 concentration was equal to the
initial one, the saturation adsorption was achieved. Similarly, a blank
Fig. 5. CO2 adsorption curve of bulk IL [P4444][2-Op] bubbled with pure CO2 at experiment had been performed by switching the mixture into the
50.0 °C. empty sample tube before the adsorption measurement of the ionogel.
The CO2 adsorption capacities (q, mmol CO2/(g ionogel) of the NS and
ionogel at a certain time (t, s) were calculated depending on the below
equation,

284
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 6. Diagram of the CO2 adsorption setup operating at ordinary pressure.

t
Fig. 8. FTIR spectra of samples (a) MS; (b) PMS-5; (c) PMS-30; (d) IL [P4444][2-
q= ∫0 Q (cin − ceff ) dt /(mvm) Op]; (e) NMS-5; (f) NMS-30; and (g) IL [N4444][2-Op].

Where m (g) represents the mass of the ionogel. Q (mL/min) is the 3.2. Thermal stability of the samples
influent flow rate. Cin (vol.%) and Ceff (vol.%) are the influent and ef-
fluent CO2 concentration respectively. The vm (22.4 mL/mmol) is Enhanced thermal stabilities of ionogels are observed from their TG
standard molar volume. Herein, ionogels PMS-10 and NMS-15 were curves (Fig. 9). The supporter MS shows a weight loss of 1.0% below
chosen to investigate the cyclic stability in 10 cycles of adsorption and 300.0 °C, which due to the removal of adsorbed H2O (Fig. 9A(a)). A
regeneration test. The process of regeneration was conducted in the further weight loss of 1.2% is observed above 300.0 °C, deriving from
pure N2 flow of 98.5 mL/min at 120.0 °C for 1.0 h. the removal of partial hydroxyl (Fig. 9A(a)). Only a weight loss smaller
than 1.0% is observed for the IL [P4444][2-Op] below 100.0 °C
(Fig. 9A(b)), coinciding with its physical water content measured on
3. Results and discussion AKF-3 Karl Fischer coulometer. An obvious weight loss up to 99.0% is
observed for it in the range from 100.0 to 350.0 °C, which can be mainly
3.1. FTIR of ILs and ionogels ascribed to its bulk decomposition (Fig. 9A(b)). Accordingly, it starts to
decompose at about 200.0 °C. While the ionogels exhibit different de-
The structures of pyridine-containing anion functionalized ILs composition behaviors. Weight losses of the ionogel PMS-5 mainly
[P4444][2-Op] and [N4444][2-Op] are simulated as in Fig. 7. Their di- occur above 400.0 °C (Fig. 9A(c)), which obviously higher than the
mensions are about 1.1 × 1.1 × 1.4 nm and suitable for loading in onset decomposition temperature of the pure IL. The results reveal that
mesoporous matrix. Here, the structure of pyridine-containing anion the thermal stability of the infused IL is enhanced by about 200.0 °C,
functionalized IL [P4444][2-Op] is confirmed by comparing its FTIR similar to that of ILs loaded in different matrices [35]. The result is
pattern. The bands of 1075.4, 800.7, and 453.4 cm–1 are observed in the different from the decreased stability of other IL on the similar matrices
supporter and ionogel (Fig. 8(a)–(c), (e), (f)) can be ascribed to the of silica [35]. They can be attributed to significant interfacial layering
vibration of bonds SieOH, SieO and SieOeSi, respectively. Some and stiffening of the immobilized IL in the vicinity of the supporter
bands at around 2936.0 cm–1 ascribed to stretching vibration of eCH3 surface as well as the lower self-diffusivity at low IL loading caused by
group also present in the ionogels and ILs (Fig. 8(b)–(g)) and derive the nanospace confinement effect [36]. The weight loss occurred at
from cation [P4444]+ or [N4444]+ [35]. Reasonably, their band in- above 400.0 °C is 60.8% of the total weight loss of the ionogel PMS-5.
tensities of high loading ionogels PMS-30 and NMS-30 (Fig. 8(c), (f)) The result implies that the IL mainly dots the mesopore surface of the
are stronger than those of low loading ionogels PMS-5 and NMS-5 supporter, in which the pore wall-IL interaction slows down its dy-
(Fig. 8(b), (e)) but weaker than the bulk ILs (Fig. 8(d), (g)). namics [35]. The ionogels PMS-1 and PMS-3 also show the main weight
losses above 400.0 °C (Fig. 10), which are similar to the results of the
ionogel PMS-5. With further infusion up to 15%, two weight losses are
observed for ionogels PMS-10 and PMS-15 (Fig. 9A(d), (e) and B(d), (e))
in the ranges of 200.0–420.0 °C and 430.0–580.0 °C, which correspond
to the decomposition of aggregated IL and nanoconfined one in ionogels
respectively. The weight loss occurred above 430.0 °C is only 25.1%
and 22.1%, separately. The results imply that the IL locates in two
different fashions and partial IL prefers to dot the mesoporous surface
and then aggregate into nanoclusters, showing high and low onset de-
composition temperature separately. The properties of IL molecules in
the nanoclusters are mainly controlled by the fluid-fluid interaction and
insignificant effect of the substrate-fluid interaction [35]. As a result,
the thermal stability in this layer is as low as the bulk IL. Some results
on imidazolium-based ionogels even show poorer thermal stability than
the free IL [37]. Interestingly, with further loading of the IL up to 30%,
Fig. 7. Optimized structure and dimensions of ILs [P4444][2-Op] (Left), and
[N4444][2-Op] (Right).
weight loss of 29.38% is mainly observed for PMS-30 in the range of

285
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 9. TG (A) and DTG (B) curves of samples: (a) MS; (b) IL [P4444][2-Op]; (c) PMS-5; (d) PMS-10; (e) PMS-15; and (f) PMS-30.

Fig. 11A(b), its onset decomposition temperature is only about


120.0 °C, which is obviously lower than the IL [P4444][2-Op]. All the
ionogels NMS-5, NMS-10, NMS-15, NMS-30 show two weight losses in
the range of 150.0–300.0 °C and 390.0–620.0 °C (Fig. 11A(c)–(f) and B
(c)–(f)), corresponding to the decomposition of aggregated IL and na-
noconfined one respectively. At the same loading amount, the results
are different from the ionogels PMS-w. In case of the ionogel NMS-5, the
IL shows two different statuses, implying lowly dispersion due to its
high viscosity (Fig. 11A(c) and B(c)). The weight loss occurred at high
temperature is obviously larger than that at low one. The result in-
dicates that IL [N4444][2-Op] prefers to dot mesopore surface of the
supporter [35]. With further infusion of the IL up to 30%, more and
more weight losses are observed for ionogels NMS-10, NMS-15, and
NMS-30 (Fig. 11A(d)–(f) and B(d)–(f)) in 150.0–300.0 °C, which is si-
milar to the IL itself (Fig. 11A(b)). Their weight losses in the range of
430–580 °C are 3.75%, 3.78% and 5.17% (Fig. 11A(d)–(f)), respec-
tively. The results reveal that most IL gradually aggregates into na-
noclusters due to the increasingly loaded amount. As a result, the
Fig. 10. TG curves recorded in N2 of IL [P4444][2-Op], PMS-3, and PMS-1. thermal stability is as low as the bulk IL in that the disappearance of
strong interfacial layering [38].
200.0–420.0 °C close to that for free IL. Only about weight loss of 2.25%
is observed above 420.0 °C. The results imply that most of the IL is 3.3. Porosity evolution of the ionogels
immobilized in a similar aggregation status such as nanoclusters, in
which the strong interfacial layering exponentially decay into the bulk The morphology of ionogels including PMS-5, PMS-10, and PMS-15
IL [38]. The results also well agree with their lowering hysteresis loops are also observed for comparing with the supporter. SEM images of all
of N2 sorption isotherms below and other reported high loading iono- the taken samples show irregular powder just as the support MS
gels [39,40]. (Fig. 12). The particle size is around 10–80 μm. It is noted that the large
Although the IL [N4444][2-Op] shows a high viscosity up to particle is loosely stacked from small ones.
982.0 mPa s, its thermal stability is very poor. Just as seen in No obvious diffraction peak can be found in XRD patterns of the

Fig. 11. TG (A) and DTG (B) curves of samples: (a) MS; (b) IL [N4444][2-Op]; (c) NMS-5; (d) NMS-10; (e) NMS-15; and (f) NMS-30.

286
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 12. SEM images (scale bar = 10 μm) of samples: (A) mesoporous slica MS, (B) PMS-5, (C) PMS-10, and (D) PMS-15.

Fig. 13. Low angle XRD patterns of the supporter and ionogels PMS (A) and NMS (B).

samples (Fig. 13A–B), indicating a disordered mesostructure. The re- (1.1 × 1.1 × 1.4 nm) of the IL [P4444][2-Op] into consideration (Fig. 7),
sults imply that the addition of TMB interrupts the formation of ordered 0.10 g IL can ideally cover the maximum area of 265.0 m2 in a mono-
mesostructure. Correspondingly, all the obtained ionogels also display molecular layer style. For the ionogel PMS-10, the 0.90 g MS with total
similar structure (Fig. 13A–B). Their N2 adsorption-desorption iso- surface area of 736.7 m2 is enough for tiling the loaded IL into a
therms and corresponding pore size distributions are summarized in monomolecular layer.
Fig. 14. The support MS shows a type IV isotherm with a hysteresis loop Accordingly, the main pore diameter evolves from the 13.5 to
at relative pressures P/P0 = 0.45–0.90 (Fig. 14A(a)), indicating the around 10.5 nm, coinciding with one IL monomolecular layer lying in
presence of mesopores [41]. Its specific surface area is calculated about mesopores. Thus, it is reasonable that the IL [P4444][2-Op] prefers to
818.5 m2/g (Table 1). Its pore size distribution is broad and mainly dot the mesopores inside and produce as large interface as possible. The
centers at 13.5 nm (Fig. 14B(a)), displaying the presence of large me- results echo on the interfacial layering on the substrate [38]. The hys-
sopore. teresis loops of PMS-5, PMS-10, PMS-15, and PMS-30 gradually shrink
Although the isotherms of the ionogels PMS-5, PMS-10, PMS-15, (Fig. 14A(b)–(e)), implying that a part of mesopores is fulfilled by the
and PMS-30 still keep type IV curves with obvious hysteresis loops IL. The ionogels based on the IL [N4444][2-Op] show a similar change
(Fig. 14A(b)–(e)), their N2 adsorption capacities gradually decrease. As trend on their specific surface area, pore size and pore volume. Briefly,
a result, the specific surface area progressively decreases to 397.5 m2/g the isotherms of the ionogels NMS-5, NMS-10, NMS-15, and NMS-30
for the ionogel PMS-30 (Table 1). Their pore volumes lower from 1.81 keep type IV curves with well-resolved hysteresis loops
to 0.88 cm3/g. The ionogels including PMS-5, PMS-10, PMS-15, and (Fig. 14C(b)–(e)). Their adsorption capacities gradually decrease due to
PMS-30 show similar pore size distributions centering around 10.5 nm the increasing introduction of the IL [N4444][2-Op]. Accordingly, their
(Fig. 14B(b)–(e) and Table 1). Taken the simulated dimensions specific surface area progressively decreases from 818.5 to 492.8 m2/g

287
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 14. N2 sorption isotherms (A) and pore size distributions (B) of samples: (a) MS; (b) PMS-5; (c) PMS-10; (d) PMS-15; and (e) PMS-30. N2 sorption isotherms (C)
and pore size distributions (D) of samples: (a) MS; (b) NMS-5; (c) NMS-10; (d) NMS-15; and (e) NMS-30.

Table 1 Because of the limited sample loading, it is difficult to observe outlet


Pore properties of the supporter MS and ionogels PMS-w and NMS-w. concentration of zero point for the discussed ionogel. According to the
Sample Specific surface area Pore diameter Pore volume
curves, the adsorption capacities of samples are calculated and sum-
(m2 g–1) (nm) (cm3 g–1) marized in Fig. 16A. The supporter MS show an adsorption capacity of
0.696 mmol CO2/(g ionogel) (Fig. 16A(a)). After loading 5% IL [P4444]
MS 818.5 13.5 1.81 [2-Op], an enhanced adsorption capacity of 1.369 mmol CO2/(g io-
PMS-5 754.8 10.1 1.61
PMS-10 620.5 11.2 1.41
nogel) (5.01 mol CO2/(mol IL)) is observed for the ionogel PMS-5
PMS-15 580.5 11.1 1.28 (Fig. 16A(b)), which is 1.967 times higher than the supporter
PMS-30 397.5 10.0 0.88 (Fig. 16A(a)). It takes about 13.0 min for the ionogel PMS-5 to achieve
NMS-5 747.2 10.1 1.58 the adsorption balance. The results indicate that the highly dispersed IL
NMS-10 719.1 11.1 1.49
brings its functional groups into full playing on the CO2 adsorption.
NMS-15 664.0 11.2 1.44
NMS-30 492.8 11.5 1.25 With further loaded IL up to 10%, the obtained ionogel PMS-10 shows
the highest CO2 adsorption capacity of 1.678 mmol CO2/(g ionogel)
(3.72 mol CO2/(mol IL)) among all the ionogels PMS-w (Fig. 16A(c)). It
(Table 1). Their pore volumes change from 1.81 to 1.24 cm3/g. All the is noteworthy to point out that the ionogel PMS-10 shows higher ca-
ionogels show similar pore size distributions (Fig. 14D(b)–(e) and pacity at 50.0 °C than that (0.905 mmol CO2/(g ionogel)) of silica im-
Table 1). All the results confirm that the IL [N4444][2-Op] firstly lines pregnated with 50% [P66614][2-Op] at 19.0 °C [30]. It is fairly compe-
the mesopores and subsequently fills the part of pores too. titive with newly reported ionogels with high loading up to 50%
[29,30,39,40]. Reasonably, it takes only about 7.5 min for PMS-10 to
reach 90% saturation adsorption, which is fairly shorter than 300.0 min
3.4. Rapid CO2 adsorption in the ionogels
required for other reported ionogels and neat ILs. The mean adsorption
rate of the ionogel PMS-10 is slightly lower than that of the sample
It takes about 70 s to achieve the lowest CO2 outlet concentration
PMS-5. Possible pore blocking by aggregated ionic liquid leads to mass
for ionogels PMS-w as seen from breakthrough curves (Fig. 15A).

288
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 15. Breakthrough curves recorded at 50.0 °C of samples.

Fig. 16. CO2 adsorption curves recorded at 50 °C of ionogels PMS-w (A): (a) MS; (b) PMS-5; (c) PMS-10; (d) PMS-15; and (e) PMS-30. Ionogels NMS-w (B): (a) MS; (b)
NMS-5; (c) NMS-10; (d) NMS-15; and (e) NMS-30.

transport limitation and low adsorption rate in PMS-10. The results It takes about 11.0 min for NMS-15 to reach the saturation adsorption,
coincide with the above results from TG analysis. With further loaded IL which is still shorter than the time required for other reported ionogels
up to 30%, the CO2 adsorption capacity is gradually recorded at 1.478, [30,40]. With further loaded IL up to 30%, the CO2 adsorption of the
1.239 mmol CO2/(g ionogel) (2.09, 0.886 mol CO2/(mol IL)) for sample ionogel NMS-30 is saturated at 0.821 mmol CO2/(g ionogel) (0.374 mol
PMS-15 and PMS-30, respectively (Fig. 16A(d) and (e)). Similar change CO2/(mol IL)) (Fig. 16B(e)). In the studies of mesoporous silica SBA-15
is also observed in the Arg/PSS-PMMA sorbents [25]. The results imply impregnated with IL of 1-ethyl-3-methylimidazolium tri[bis(tri-
that less and less active site is accessible for CO2 molecule due to the fluoromethylsulfonyl)imide]zincate and nano- porous PMMA supported
formation of more and more IL nanoclusters. IL [EMIM][Gly], analogous trends are also observed [26,42].
In case of the ionogels containing IL [N4444][2-Op], they spend
different time to achieve the lowest CO2 outlet concentration (Fig. 15B).
Their adsorption capacities are also calculated and summarized in 3.5. Fabrication mechanism of the nanostructured ionogels and their CO2
Fig. 16B. After loading 5.0% IL, an adsorption capacity of 0.905 mmol adsorption
CO2/(g ionogel) (1.64 mol CO2/(mol IL)) is observed for the ionogel
NMS-5 (Fig. 16B(b)). It takes about 9.5 min to achieve the saturation Basing on the above results, we propose a fabrication mechanism of
adsorption. With the loaded IL up to 10.0%, the sample NMS-10 shows the ionogel by selecting the ionogel PMS-w as a model (Fig. 17). The
the adsorption capacity of 0.991 mmol CO2/(g ionogel) (1.23 mol CO2/ supporter MS with large pore is prepared by mixing pore-expanding
(mol IL)) (Fig. 16B(c)). The highest adsorption capacity of 1.081 mmol agent and structure-directing one in the synthetic system. Pyridine-
CO2/(g ionogel) (1.10 mol CO2/(mol IL)) is observed for the ionogel containing anion-functionalized IL [P4444][2-Op] is prepared and fur-
NMS-15 among all the ionogels (Fig. 16B(d)). Although the loaded ther infused into the MS. Based on the calculation, the cation [P4444]+
amount of IL in the supporter is tripled, the adsorption capacity of NMS- prefers to line on the vicinity of the silica surface due to its lower
15 does not increase proportionally compared with the ionogel NMS-5. bounding energy of –41.8 kJ/mol than that (–37.7 kJ/mol) of the anion
[2-Op]– close to silica surface and that (–22.3 kJ/mol) of the IL pair in a

289
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 17. Fabrication mechanism of the ionogel by loading IL [P4444][2-Op] onto


large mesoporous silica.

staggered style [32]. The simulation on the surface electrostatic po-


tential also indicates the preference of the cation [P4444]+ close to the
surface of silica [32]. The lowly loaded cation [P4444]+ is closely flatten
Fig. 18. CO2 adsorption capacity of three batches of ionogels PMS-w (Black,
to be more compatible with rough pore walls due to the strong inter-
left) and NMS-w (Red, right) at 50.0 °C.
facial layering [38,43]. Meantime, the anion [2-Op]– is well accessible
for the guest molecule CO2. The appearance of two new peaks in the 13C
NMR of IL [P4444][2-Op] after capturing CO2 also confirms the CO2
adsorption (Fig. 5(b)). The low loaded ionogel PMS-5 may present
sparse but well exposed IL monomolecular islands, allowing the oc-
currence of non-stoichiometric CO2 adsorption. As a result, the de-
clining in pore diameter well echoes the monolayer coating of IL. With
the ascent of IL lined up to 10%, more IL monomolecular islands be-
come bumpy but scarcely IL pairs or nanoclusters start to form in the
large mesopores (Fig. 17). The former shows an enhanced thermal
stability but the latter shows poor one just as the bulk IL based on the
TG results. The non-stoichiometric CO2 adsorption is still observed for
the ionogel PMS-10. Reasonably, long time should be left for CO2 dif-
fusing into the IL nanoclusters. Finally, the CO2 adsorption capacity
increases in sequence with the increasing IL loading but a relatively low
rate. The ionogel PMS-10 shows the attractive CO2 adsorption capacity
up to 1.678 mmol CO2/(g ionogel) (3.72 mol CO2/(mol IL)). Many ILs
for the separation of CO2 from N2 are too inefficient to be economically
viable. Here, the rapid adsorption in the mixture of CO2/N2 can be
achieved on the ionogel PMS-10 with sound stability. Due to the
thronging IL up to 15%, one part of the IL still prefers to flat as islands
Fig. 19. Cyclic CO2 adsorption capacity recorded at 50.0 °C of ionogels: (a)
and another part states as nanoclusters, occupying pore volume and
PMS-10; and (b) NMS-15.
showing same properties as the bulk IL (Fig. 17). The ionogel PMS-15
shows two distinguishingly different onset decomposition temperatures
and a decreasing CO2 adsorption capacity. Further loaded IL up to 30% PMS-10 in each cycle is summarized in Fig. 19(a). The capacity of the
aggregates into IL pairs even clusters (Fig. 17), which fill part of pores ionogel PMS-10 fluctuates from 1.678 to 1.511 mmol CO2/(g ionogel)
and show the same thermal stability to the bulk IL. It is not surprising (3.13 mol CO2/(mol IL)) at the tenth cycle. About adsorption-capacity
that low surface area and pore volume only allow a poor CO2 adsorp- retention of 90% shows a good cyclic stability and sorption reversi-
tion. In a summary, loading-dependent status of the IL companied with bility. The cyclic stability can be attributed to the modified physico-
dramatically modified properties figure out that low content supported chemical property of [P4444][2-Op] profiting from nano-confinement
IL steadily show rapid CO2 sorption at a constant pressure. effect, such as strong interfacial layering, higher viscosity, and lower
self-diffusion. All the properties cooperatively make the low loaded IL
well anchor in the nanoporous wall and possess good endurance to
3.6. Cyclic stability of CO2 adsorption tough regeneration condition. The adsorption capacity of the ionogel
NMS-15 floats from 1.081 to 0.961 mmol CO2/(g ionogel) (0.829 mol
CO2 adsorption capacities from three batches of ionogels are sum- CO2/(mol IL)) (Fig. 19(b)). It capacity retention of about 89% is ob-
marized for further comparison (Fig. 18). All the ionogels exhibit steady served, implying a good cyclic stability. It can be attributed to itself
CO2 adsorption capacity at 50.0 °C and CO2 partial pressure of 14.9%. strong hydrogen-bond interaction and high viscosity.
With the increased IL, CO2 adsorption capacities of two types of iono-
gels gradually go up and then go down, which are closely related to the
dispersion status of the IL in MS. In total, the ionogels PMS-10 and 3.7. Influences of CO2 partial pressure on the adsorption capacity and rate
NMS-15 show the attractive adsorption capacity among them. Excellent
cyclic stability is often more attractive than high adsorption capacity Accordingly, a typical coal-fired flue gas contains 76–77% N2,
for an adsorbent in practical applications. Here, we take the ionogels 12.5–12.8% CO2, 6.2% H2O, 4.4% O2, and trace amount of CO, NOx,
PMS-10 and NMS-15 to investigate the cyclic stability by performing 10 and SO2 gas [44]. We further investigate the CO2 adsorption capacity at
cycles of CO2 adsorption. The CO2 adsorption capacity of the ionogel different partial pressure on the ionogels PMS-10 and NMS-15. In

290
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Fig. 20. CO2 adsorption capacity of ionogels PMS-10 (A) and NMS-15 (B) at 50.0 °C in gas mixture with CO2 partial pressure: (a) 0.244, (b) 0.149, (c) 0.097, and (d)
0.052.

details, an adsorption capacity of 1.716 mmol CO2/(g ionogel)


(3.85 mol CO2/(mol IL)) is observed after exposing PMS-10 at CO2
partial pressure of 0.224 (Fig. 20A(a)), which is close to that balancing
at partial pressure of 0.149 (Fig. 20A(b)). However, it takes only 400 s
to achieve the saturation adsorption, which is obviously faster than that
(1100 s) at CO2 partial pressure of 0.149. The results indicate that high
partial pressure impose more effect on the adsorption process than final
capacity. The adsorption capacities of PMS-10 are 1.508 and
1.357 mmol CO2/(g ionogel) (3.12, and 2.58 mol CO2/(mol IL)) at
partial pressures of 0.097 and 0.052, respectively (Fig. 20A(c) and (d)).
It takes relatively short time to achieve the saturation adsorption for it
at the low CO2 partial pressure. Adsorption capacity obviously shrinks
during the initial fast adsorption process, which is controlled by ex-
ternal diffusion. Lower adsorption rate is observed in the second ad-
sorption stage due to the decaying driving force for CO2 diffusion in
PMS-10.
Similarly, a saturation adsorption capacity of 1.119 mmol CO2/(g
ionogel) (1.18 mol CO2/(mol IL)) is observed for ionogel NMS-15 at
CO2 partial pressure of 0.224 (Fig. 19B(a)), which is very close to the Fig. 21. CO2 adsorption capacity of ionogels PMS-10 (a) and NMS-15 (b) at
capacity of 1.081 mmol CO2/(g ionogel) (1.10 mol CO2/(mol IL)) at 30.0, 50.0, and 70.0 °C recorded in gas mixture with CO2 partial pressure of
partial pressure of 0.149 (Fig. 20B(b)). It takes 440 s for the ionogel 0.149.
NMS-15 to achieve the saturation adsorption, which is obviously faster
than that at low CO2 partial pressure. Reasonably, the adsorption ca-
obtaining high adsorption capacity.
pacities of NMS-15 are 0.752 and 0.579 mmol CO2/(g ionogel) at lower
partial pressures of 0.097 and 0.052 (Fig. 20B(c) and (d)). Additionally,
4. Conclusions
the lowest capacity is calculated at the smallest partial pressure basing
on the equilibrium time of 590 s. In general, the ionogel NMS-15 can
Rapid CO2 adsorption is demonstrated on the mesoporous ionogels
exhibit the higher and higher CO2 adsorption capacity in shorter and
[P4444][2-Op]@MS and [N4444][2-Op]@MS, which are prepared by
shorter time with increasing CO2 partial pressure.
immobilizing ILs tetrabutylammonium 2-hydroxypyridine ([N4444][2-
Op]) and tetrabutylphosphonium 2-hydroxypyridine ([P4444][2-Op])
3.8. Influences of CO2 temperature on the adsorption capacity into large pores. It only takes 10.0 min for the ionogel NMS-15 to
achieve the saturated adsorption capacity of 1.081 mmol CO2/(g io-
Operation temperature of the flue gas is generally below 80.0 °C nogel) at 50.0 °C in a mixture with low CO2 partial pressure. It takes
[44,45]. In view of this point, the temperature dependence of final 7.5 min to achieve 90% saturated adsorption capacity of 1.678 mmol
adsorption capacity is also studied on ionogels PMS-10 and NMS-15 in CO2/(g ionogel) (3.72 mol CO2/(mol IL)) for the ionogel PMS-10, which
CO2 with partial pressure of 0.149 at the temperatures 30.0, 50.0 and is 12.76 times higher than that of the bulk IL. Its adsorption-capacity
70.0 °C. It can be seen that the ionogel PMS-10 exhibits the highest retention of around 90% is observed after performing 10 cycles sorption
adsorption capacity of 1.85 mmol CO2/(g ionogel) (4.33 mol CO2/(mol test. The results can be ascribed to the MS with large pore that benefits
IL)) at 30.0 °C and the poorest one of 1.50 mmol CO2/(g ionogel) to not only load more exposed IL but also leave necessary pathway for
(3.09 mol CO2/(mol IL)) at 70.0 °C, respectively (Fig. 21(a)). The trend CO2 diffusion. Featuring with low loading and cost, high capacity, swift
is reasonable and similar to the reported absorbent, which can be as- adsorption and good cyclic stability make the ionogel PMS-10 a robust
cribed to the weakened interaction between the CO2 and the ionogel candidate in CO2 capture from the flue gas.
PMS-10 as the temperature rising [26]. The adsorption capacities of the
ionogel NMS-15 are 1.17, 1.08 and 1.00 mmol CO2/(g ionogel) (1.30, Acknowledgements
1.10, and 0.916 mol CO2/(mol IL)) at 30.0, 50.0 and 70.0 °C
(Fig. 21(b)). All the results indicate that low temperature is helpful in The authors acknowledge the financial supports from China

291
C. Xue, et al. Journal of CO₂ Utilization 34 (2019) 282–292

Scholarship Council (No. 201506935028), Key Scientific and [24] Y. Uehara, D. Karami, N. Mahinpey, Roles of cation and anion of amino acid anion-
Technological Project of Shanxi Province (No. MD2014-09), and functionalized ionic liquids immobilized into a porous support for CO2 capture,
Energy Fuels 32 (2018) 5345–5354.
Natural Science Foundation of Shanxi Province (No. 201801D121060). [25] B. Jiang, X. Wang, M.L. Gray, Y. Duan, D. Luebke, B. Li, Development of amino acid
and amino acid-complex based solid sorbents for CO2 capture, Appl. Energy 109
References (2013) 112–118.
[26] X. Wang, N.G. Akhmedov, Y. Duan, D. Luebke, B. Li, Immobilization of amino acid
ionic liquids into nanoporous microspheres as robust sorbents for CO2 capture, J.
[1] S. Chu, Carbon capture and sequestration, Science 325 (2009) 1599. Mater. Chem. A 1 (2013) 2978–2982.
[2] J. Wang, L. Huang, R. Yang, Z. Zhang, J. Wu, Y. Gao, Q. Wang, D. O’Hare, Z. Zhong, [27] S. Lee, S.-Y. Moon, H. Kim, J.-S. Bae, E. Jeon, H.-Y. Ahn, J.-W. Park, Carbon dioxide
Recent advances in solid sorbents for CO2 capture and new development trends, absorption by hydroxyalkyl amidines impregnated into mesoporous silica: the effect
Energy Environ. Sci. 7 (2014) 3478–3518. of pore morphology and absorbent loading, RSC Adv. 4 (2014) 1543–1550.
[3] Q. Wang, J. Luo, Z. Zhong, A. Borgna, CO2 capture by solid adsorbents and their [28] I.H. Arellano, S.H. Madani, J. Huang, P. Pendleton, Carbon dioxide adsorption by
applications: current status and new trends, Energy Environ. Sci. 4 (2011) 42–55. zinc-functionalized ionic liquid impregnated into bio-templated mesoporous silica
[4] S. Li, Z. Wang, X. Yu, J. Wang, S. Wang, High-performance membranes with multi- beads, Chem. Eng. J. 283 (2016) 692–702.
permselectivity for CO2 separation, Adv. Mater. 24 (2012) 3196–3200. [29] J. Zhang, S. Zhang, K. Dong, Y. Zhang, Y. Shen, X. Lv, Supported absorption of CO2
[5] A.I. Cooper, Materials chemistry: cooperative carbon capture, Nature 519 (2015) by tetrabutylphosphonium amino acid ionic liquids, Chem. Eur. J. 12 (2006)
294–295. 4021–4026.
[6] N. MacDowell, N. Florin, A. Buchard, J. Hallett, A. Galindo, G. Jackson, [30] J. Cheng, Y. Li, L. Hu, J. Zhou, K. Cen, CO2 adsorption performance of ionic liquid
C.S. Adjiman, C.K. Williams, N. Shah, P. Fennell, An overview of CO2 capture [P66614][2-Op] loaded onto molecular sieve MCM-41 compared to pure ionic liquid
technologies, Energy Environ. Sci. 3 (2010) 1645–1669. in Biohythane/Pure CO2 atmospheres, Energy Fuels 30 (2016) 3251–3256.
[7] G.T. Rochelle, Amine scrubbing for CO2 capture, Science 325 (2009) 1652–1654. [31] G.E. Romanos, P.S. Schulz, M. Bahlmann, P. Wasserscheid, A. Sapalidis,
[8] M.E. Boot-Handford, J.C. Abanades, E.J. Anthony, M.J. Blunt, S. Brandani, N. Mac F.K. Katsaros, C.P. Athanasekou, K. Beltsios, N.K. Kanellopoulos, CO2 capture by
Dowell, J.R. Fernandez, M.-C. Ferrari, R. Gross, J.P. Hallett, R.S. Haszeldine, novel supported ionic liquid phase systems consisting of silica nanoparticles en-
P. Heptonstall, A. Lyngfelt, Z. Makuch, E. Mangano, R.T.J. Porter, capsulating amine-functionalized ionic liquids, J. Phys. Chem. C 118 (2014)
M. Pourkashanian, G.T. Rochelle, N. Shah, J.G. Yao, P.S. Fennell, Carbon capture 24437–24451.
and storage update, Energy Environ. Sci. 7 (2014) 130–189. [32] C. Xue, H. Zhu, X. Du, X. An, E. Wang, D. Duan, L. Shi, X. Hao, B. Xiao, C. Peng,
[9] L.B. Sun, Y.H. Kang, Y.Q. Shi, Y. Jiang, X.Q. Liu, Highly selective capture of the Unique allosteric effect-driven rapid adsorption of carbon dioxide in a newly de-
greenhouse gas CO2 in polymers, ACS Sustain. Chem. Eng. 3 (2015) 3077–3085. signed ionogel [P4444][2-Op]@MCM-41 with excellent cyclic stability and loading-
[10] P.-Q. Liao, X.-W. Chen, S.-Y. Liu, X.-Y. Li, Y.-T. Xu, M. Tang, Z. Rui, H. Ji, J.- dependent capacity, J. Mater. Chem. A 5 (2017) 6504–6514.
P. Zhang, X.-M. Chen, Putting an ultrahigh concentration of amine groups into a [33] H. Gao, M. Xi, X. Qi, M. Lu, T. Zhan, W. Sun, Application of a hydroxyl functio-
metal-organic framework for CO2 capture at low pressures, Chem. Sci. 7 (2016) nalized ionic liquid modified electrode for the sensitive detection of adenosine-5′-
6528–6533. monophosphate, J. Electroanal. Chem. 664 (2012) 88–93.
[11] H. Wang, B. Li, H. Wu, T.-L. Hu, Z. Yao, W. Zhou, S. Xiang, B. Chen, A flexible [34] D. Jiang, J. Gao, J. Yang, W. Su, Q. Yang, C. Li, Mesoporous ethane-silicas func-
microporous hydrogen-bonded organic framework for gas sorption and separation, tionalized with trans-(1R,2R)-diaminocyclohexane: relation between structure and
J. Am. Chem. Soc. 137 (2015) 9963–9970. catalytic properties in asymmetric transfer hydrogenation, Microporous
[12] C. Serre, S. Bourrelly, A. Vimont, N.A. Ramsahye, G. Maurin, P.L. Llewellyn, Mesoporous Mater. 105 (2007) 204–210.
M. Daturi, Y. Filinchuk, O. Leynaud, P. Barnes, G. Férey, An explanation for the very [35] M.P. Singh, R.K. Singh, S. Chandra, Ionic liquids confined in porous matrices:
large breathing effect of a metal–organic framework during CO2 adsorption, Adv. physicochemical properties and applications, Prog. Mater. Sci. 64 (2014) 73–120.
Mater. 19 (2007) 2246–2251. [36] B. Coasne, L. Viau, A. Vioux, Loading-controlled stiffening in nanoconfined ionic
[13] M. Ramdin, T.W. de Loos, T.J.H. Vlugt, State-of-the-Art of CO2 capture with ionic liquids, J. Phys. Chem. Lett. 2 (2011) 1150–1154.
liquids, Ind. Eng. Chem. Res. 51 (2012) 8149–8177. [37] A.K. Tripathi, Y.L. Verma, R.K. Singh, Thermal, electrical and structural studies on
[14] J. Zhang, C. Jia, H. Dong, J. Wang, X. Zhang, S. Zhang, A novel dual amino-func- ionic liquid confined in ordered mesoporous MCM-41, J. Mater. Chem. A 3 (2015)
tionalized cation-tethered ionic liquid for CO2 capture, Ind. Eng. Chem. Res. 52 23809–23820.
(2013) 5835–5841. [38] M. Mezger, H. Schröder, H. Reichert, S. Schramm, J.S. Okasinski, S. Schöder,
[15] S. Ye, X. Jiang, L.-W. Ruan, B. Liu, Y.-M. Wang, J.-F. Zhu, L.-G. Qiu, Post-com- V. Honkimäki, M. Deutsch, B.M. Ocko, J. Ralston, M. Rohwerder, M. Stratmann,
bustion CO2 capture with the HKUST-1 and MIL-101(Cr) metal–organic frame- H. Dosch, Molecular layering of fluorinated ionic liquids at a charged sapphire
works: adsorption, separation and regeneration investigations, Microporous (0001) surface, Science 322 (2008) 424–428.
Mesoporous Mater. 179 (2013) 191–197. [39] M.M. Wan, H.Y. Zhu, Y.Y. Li, J. Ma, S. Liu, J.H. Zhu, Novel CO2-capture derived
[16] S. Supasitmongkol, P. Styring, High CO2 solubility in ionic liquids and a tetraalk- from the basic ionic liquids orientated on mesoporous materials, ACS Appl. Mater.
ylammonium-based poly(ionic liquid), Energy Environ. Sci. 3 (2010) 1961–1972. Interfaces 6 (2014) 12947–12955.
[17] X. Luo, Y. Guo, F. Ding, H. Zhao, G. Cui, H. Li, C. Wang, Significant improvements in [40] Y. Zhou, J. Liu, M. Xiao, Y. Meng, L. Sun, Designing supported ionic liquids (ILs)
CO2 capture by pyridine-containing anion-functionalized ionic liquids through within inorganic nanosheets for CO2 capture applications, ACS Appl. Mater.
multiple-site cooperative interactions, Angew. Chem. Int. Ed. 53 (2014) Interfaces 8 (2016) 5547–5555.
7053–7057. [41] C. Xue, B. Tu, D. Zhao, Evaporation-induced coating and self-assembly of ordered
[18] L.C. Tome, I.M. Marrucho, Ionic liquid-based materials: a platform to design en- mesoporous carbon-silica composite monoliths with macroporous architecture on
gineered CO2 separation membranes, Chem. Soc. Rev. 45 (2016) 2785–2824. polyurethane foams, Adv. Funct. Mater. 18 (2008) 3914–3921.
[19] Y. Yu, J. Mai, L. Huang, L. Wang, X. Li, Ship in a bottle synthesis of ionic liquids in [42] I.H. Arellano, J. Huang, P. Pendleton, Synergistic enhancement of CO2 uptake in
NaY supercages for CO2 capture, RSC Adv. 4 (2014) 12756–12762. highly ordered mesoporous silica-supported zinc-functionalized ionic liquid sor-
[20] A.I. Horowitz, M.J. Panzer, Poly(dimethylsiloxane)-supported ionogels with a high bents, Chem. Eng. J. 281 (2015) 119–125.
ionic liquid loading, Angew. Chem. Int. Ed. 53 (2014) 9780–9783. [43] Y. Arasaki, K. Takatsuka, Optical conversion of conical intersection to avoided
[21] M.-A. Néouze, J.L. Bideau, P. Gaveau, S. Bellayer, A. Vioux, Ionogels, new materials crossing, J. Chem. Soc. Faraday Trans. 12 (2010) 1239–1242.
arising from the confinement of ionic liquids within silica-derived networks, Chem. [44] N. Du, H.B. Park, M.M. Dal-Cin, M.D. Guiver, Advances in high permeability
Mater. 18 (2006) 3931–3936. polymeric membrane materials for CO2 separations, Energy Environ. Sci. 5 (2012)
[22] L.J. Lozano, C. Godínez, A.P. de los Ríos, F.J. Hernández-Fernández, S. Sánchez- 7306–7322.
Segado, F.J. Alguacil, Recent advances in supported ionic liquid membrane tech- [45] S.D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A.G. Russell, Q. Wei, M. Fan,
nology, J. Memb. Sci. 376 (2011) 1–14. Review of recent advances in carbon dioxide separation and capture, RSC Adv. 3
[23] B. Marco, E. Alessandro, L. Amedeo, T. Giorgio, M. Fabio, T. Rosa, Post-combustion (2013) 22739–22773.
CO2 capture: on the potentiality of amino acid ionic liquid as modifying agent of
mesoporous solids, Fuel 218 (2018) 155–161.

292

Das könnte Ihnen auch gefallen