You are on page 1of 8

Adsorption (2016) 22:581–588

DOI 10.1007/s10450-015-9734-0

Equilibrium isotherms and isosteric heat for CO2 adsorption


on nanoporous carbons from polymers
Jerzy Choma1 • Kamila Stachurska1 • Michal Marszewski2 • Mietek Jaroniec2

Received: 27 September 2015 / Revised: 24 November 2015 / Accepted: 26 November 2015 / Published online: 8 December 2015
Ó Springer Science+Business Media New York 2015

Abstract Four nanoporous carbons obtained from dif- make them viable candidates for CO2 capture, and for other
ferent polymers: polypyrrole, polyvinylidene fluoride, sul- adsorption and environmental-related applications.
fonated styrene–divinylbenzene resin, and phenol–
formaldehyde resin, were investigated as potential adsor- Keywords Nanoporous carbons  Polymer-derived
bents for carbon dioxide. CO2 adsorption isotherms mea- carbons  CO2 adsorption  Isosteric heat of adsorption 
sured at eight temperatures between 0 and 60 °C were used Microporosity
to study adsorption properties of these polymer-derived
carbons, especially CO2 uptakes at ambient pressure and
different temperatures, working capacity, and isosteric heat 1 Introduction
of adsorption. The specific surface areas and the volumes
of micropores and ultramicropores estimated for these The newest report of ‘‘The International Panel on Climate
materials by using the density functional theory-based Change’’ clearly indicates the ongoing and deepening cli-
software for pore size analysis ranged from 840 to mate change (IPCC, Climate Change 2013). The report
1990 m2 g-1, from 0.22 to 1.47 cm3 g-1, and from 0.18 to states the anthropogenic greenhouse gas emissions con-
0.64 cm3 g-1, respectively. The observed differences in tribute mostly to the climate change, and the resulting
the nanoporosity of these carbons had a pronounced effect changes already have a negative impact on human health,
on the CO2 adsorption properties. The highest CO2 agriculture, water resources, and ecosystems on land and
uptakes, 6.92 mmol g-1 (0 °C, 1 atm) and 1.89 mmol g-1 ocean. The main source of greenhouse gases—primarily
(60 °C, 1 atm), were obtained for the polypyrrole-derived CO2, water vapor, CH4 and N2O (Piacentini and Mujumdar
activated carbon prepared through a single carbonization- 2009)—is burning of fossil fuels, in industrial processes or
KOH activation step. The working capacity for this otherwise. The IPCC stressed the importance of reduction
adsorbent was estimated to be 3.70 mmol g-1. Depending of greenhouse gas emissions, particularly the anthro-
on the adsorbent, the CO2 isosteric heats of adsorption pogenic CO2 emissions, and proposed carbon capture and
varied from 32.9 to 16.3 kJ mol-1 in 0–2.5 mmol g-1 storage (CCS) as a viable solution. Adsorption on nano-
range. Overall, the carbons studied showed well-developed porous solids is one of the possible implementations of
microporosity and exceptional CO2 adsorption, which CCS.
So far various materials for CO2 adsorption and CCS
were explored such as nanoporous carbons (Ludwinowicz
and Jaroniec 2015a; Seredych et al. 2014), modified porous
& Mietek Jaroniec silica (Heydari-Gorji et al. 2011) and organosilica (De
jaroniec@kent.edu
Canck et al. 2013), zeolites (Su and Lu 2012), porous
1
Institute of Chemistry, Military Technical Academy, polymers (Patel et al. 2012), and organometallic networks
00-908 Warsaw, Poland (Liu et al. 2012). Among these materials, activated carbons
2
Department of Chemistry and Biochemistry, Kent State commonly have high surface areas, large pore volumes,
University, Kent, OH 44242, USA and proper pore sizes; all beneficial for CO2 adsorption. In

123
582 Adsorption (2016) 22:581–588

addition, carbons can be prepared in variety of morpholo- polymerization of pyrrole (3.8 g) in the presence of FeCl3
gies, tailorable surface chemistry, and high chemical (wa- (200 mL of 0.5 M FeCl3 aqueous solution) under stirring
ter, acids and bases) and thermal stability (Liu et al. 2015). for 2 h at room temperature; afterwards, the product was
Activated carbons are commonly prepared from natural filtered, washed with distilled water, and dried at 80 °C for
precursors, such as coal, lignite, peat, wood, fruit pits, coal 12 h. The resulting polymer was mixed with KOH (1:4
tar, and various biomasses (Jankowska et al. 1991; Marsh C:KOH weight ratio) and heated at 700 °C for 2 h using
and Rodriguez-Reinoso 2006). 3 °C min-1 temperature ramp in nitrogen flow
Alternatively, synthetic wastes can be used as carbon (40 dm3 h-1). The resulting activated carbon was washed
precursors as well. For instance, polymers or polymeric with a 1 M HCl aqueous solution and distilled water, and
wastes used as carbon precursors include: polyethylene (PE), dried at 110 °C for 12 h.
polypropylene (PP), polystyrene (PS), poly(vinyl chloride)
(PVC), poly(ethylene terephtalate) (PET), poly(methyl 2.1.2 Activated carbon from polyvinylidene fluoride (AC-
metacrylate) (PMMA), polyacrylonitrile (PAN), and more PVDF)
(Mishra et al. 2003; Arienillas et al. 2005; Migahead et al.
2010; Gupta et al. 2014; Choma et al. 2014a). These, and Another activated carbon studied, AC-PVDF, was reported
other polymers, are used every day for production of com- by Choma et al. (2015). In this case, polyvinylidene fluo-
mon objects, e.g., building components, household appli- ride granules were carbonized at 450 °C for 30 min in
ances, textiles, CDs and DVDs, and computer equipment. flowing nitrogen (heating rate = 5 °C min-1 and nitrogen
Carbons prepared by carbonization of polymers typically flow = 100 dm3 h-1). Then, the resulting carbon was
require subsequent activation to develop their surface area mixed with KOH (1:5 C:KOH weight ratio) and activated
and porosity to be viable adsorbents. Activation is typically at 700 °C for 1 h using the same heating rate and nitrogen
classified as either physical or chemical, with the first flow as in the case of carbonization. Finally, the activated
including activation with CO2 or water vapor (Górka and carbon was washed with a 35 wt% HCl aqueous solution
Jaroniec 2011), and the second, activation with KOH (de and distilled water, and dried at 100 °C for 24 h.
Souza et al. 2013) or ZnCl2 (Ludwinowicz and Jaroniec
2015b). Wickramaratne and Jaroniec (2013) found that the 2.1.3 Activated carbon from sulfonated styrene–
CO2 activation of carbon spheres prepared by Stöber method divinylbenzene resin (AC-SDVB)
resulted in the development of microporosity and high CO2
uptake of 8.05 mmol g-1 (0 °C, 1 atm). Activated carbons The third activated carbon studied, AC-SVDB, was pre-
are effective CO2 adsorbents mainly due to the large amount pared according to the recipe of Choma et al. (2014b) using
of ultramicropores (micropores with size \1 nm), which a commercially available sulfonated styrene–divinylben-
were shown to enhance CO2 adsorption (Lee and Park 2013; zene resin Amberjet 1200H (Rohm and Haas, Philadelphia,
Liu et al. 2015). USA). The resin beads (d & 5 mm; 120 g) were dried at
In this article, we examine four carbon adsorbents pre- 100 °C for 3 h, immersed in 300 cm3 of 50 wt%
pared by carbonizing and activating different polymers, orthophosphoric acid, and evaporated in a rotary evapora-
study their structure and CO2 adsorption properties at dif- tor at 180 °C. Next, the cooled down sample was car-
ferent temperatures ranging from 0 to 60 °C. This study bonized at 350 °C for 90 min under static conditions (i.e.,
aims to explore the opportunities in the development of no gas flow), followed by activation with KOH (1:4
microporosity in the polymer-derived carbons for enhanc- C:KOH weight ratio) at 700 °C, washing and drying as
ing CO2 uptake at ambient conditions. reported by Choma et al. (2014b).

2.1.4 Ordered mesoporous carbon from phenol–


2 Experimental formaldehyde resin (OMC-PF)

2.1 Materials OMC-PF ordered mesoporous carbon was prepared by


soft-templating procedure reported by Choma et al. (2012).
2.1.1 Activated carbon from polypyrrole (AC-PPy) The XRD patterns and TEM images for phenolic resin-
based OMCs, confirming their mesostructural ordering, are
One of the samples studied, namely polypyrrole-derived already available in literature (de Souza et al. 2013; Wang
activated carbon (AC-PPy), was reported by Choma et al. et al. 2008). The synthesis was carried out under acidic
(2015). This carbon (AC-PPy) was obtained by a single- (HCl) conditions using resorcinol and formaldehyde as the
step carbonization-KOH activation of a self-made poly- carbon precursors, and Pluronic F127 triblock copolymer
pyrrole (PPy). First, polypyrrole was prepared by (polyethylene oxide-polypropylene oxide-polyethylene

123
Adsorption (2016) 22:581–588 583

oxide) as the structure-directing agent. First, 2.5 g of maxima of the first and second PSD peaks, respectively.
resorcinol and 2.5 g of Pluronic F127 block copolymer Microporosity was calculated as the ratio of V2 to Vt and
were dissolved in a mixture of 11.9 mL of ethanol and expressed in %. Isosteric heats of CO2 adsorption (qst) were
6.6 mL of distilled water. After complete dissolution, calculated based on the CO2 adsorption isotherms at 0, 10,
2.2 mL of HCl (35–38 wt% aqueous solution) was added 20, 25, 30, 40, 50, and 60 °C using the Clausius–Clapeyron
and the mixture was stirred for 30 min. Then, 2.5 mL of equation:
formaldehyde (37 wt% aqueous solution) was added and  
o ln p
the stirring was continued until the solution separated into qst ¼ RT2
oT a
two layers. The bottom organic layer was collected,
transferred to a Petri dish, and dried at 100 °C for 24 h. where R is the universal gas constant, T is the adsorption
The resulting organic–organic composite was carbonized temperature, p is the equilibrium pressure, and a is the
in nitrogen flow (20 dm3 h-1) using the following tem- amount adsorbed (Myers 2002).
perature program: first, temperature was ramped to 180 °C
using 2 °C min-1 and dwelled for 5 h, then, temperature
was ramped to 400 °C using 2 °C min-1, and immediately 3 Results and discussion
(no dwell), temperature was ramped to 850 °C using
5 °C min-1 and dwelled for 2 h. The procedure yielded The CO2 adsorption properties of four selected carbon
&1 g of the carbon product. adsorbents derived from different polymers were investi-
gated to explore the opportunities in the development of
2.2 Measurements microporosity in the polymer-derived carbons for achiev-
ing high CO2 uptakes at ambient pressure and different
Nitrogen adsorption–desorption isotherms at -196 °C and temperatures between 0 and 60 °C. A series of eight CO2
carbon dioxide adsorption isotherms at 0, 10, 20, 25, 30, adsorption isotherms was measured for each of the four
40, 50, and 60 °C were measured on an ASAP 2020 surface carbon studied. The corresponding nitrogen adsorption
area and porosity analyzer manufactured by Micromeritics isotherms and PSDs are presented in Fig. 1. The isotherm
Instrument Corp., Norcross, GA, USA. Each sample was obtained for AC-PPy shows type I(a) according to the new
degassed at 200 °C for 2 h in vacuum before every IUPAC classification (Thommes et al. 2015), indicating the
measurement. presence of small micropores. On the other hand, the iso-
therm recorded for AC-PVDF is type I(b), which also
2.3 Calculations indicates microporous structure, but one with broad pore
size distribution and possibly small mesopores. In contrast,
The apparent specific surface area (SBET) was determined AC-SDVB and OMC-PF both have isotherms being a
based on the low-temperature nitrogen adsorption by the combination of type I and IV (Kruk and Jaroniec 2001).
Brunauer–Emmett–Teller method (Brunauer et al. 1938) in The PSD curves support these observations: AC-PPy has
the relative pressure range 0.05–0.20 using the nitrogen exclusively pores \1 nm, AC-PVDF has a bimodal PSD
cross-section area of 0.162 nm2. Total pore volume (Vt) spanning from small micropores up to & 4 nm, and AC-
was calculated by converting the volume of nitrogen SDVB and OMC-PF have both micropores and mesopores
adsorbed at the relative pressure &0.99 to the volume of present, although of different sizes and quantities.
liquid nitrogen at the experiment’s conditions. Pore size Table 1 lists the structural parameters calculated based
distributions (PSD) and specific surface areas (SDFT) were on nitrogen adsorption isotherms for all carbons studied.
calculated based on the low-temperature nitrogen adsorp- These carbons feature very large specific surface areas,
tion using the non-local density functional theory method SDFT, reaching 2000 m2 g-1 in all instances and the
(NLDFT) for carbon slit-shaped pores with energetic apparent SBET surface areas from 1810 to 2920 m2 g-1,
heterogeneity and geometrical corrugation developed by except OMC-PF, which had a smaller surface area,
Jagiello and Olivier (2013a, b) and implemented in SDFT = 840 m2 g-1 and SBET = 660 m2 g-1. Similarly,
SAIEUS software by Jagiello. Volumes of micropores and the total pore volumes are large for these carbons in the
small mesopores (V2) were calculated by integration of the range of 0.87–1.64 cm3 g-1, and slightly smaller for OMC-
PSD curves under the first and second peak in the range PF, Vt = 0.65 cm3 g-1. All carbons possess ultramicrop-
from zero to the endpoint of the second peak. Volumes of ores (pores \ 1 nm, labeled as V1) ranging from 0.18 to
small micropores (V1) were calculated by integration of the 0.64 cm3 g-1 depending on the polymer used. The average
PSD curves under the first peak in the range from zero to size of these pores (w1) varies from 0.6 to 0.9 nm, as
the PSD minimum after the first peak. The sizes of small indicated by the position of the first PSD peak for each
(w1) and large (w2) micropores were estimated at the carbon sample. The carbons possess larger micropores as

123
584 Adsorption (2016) 22:581–588

Fig. 1 Nitrogen adsorption–desorption isotherms at -196 °C (left from polypyrrole, polyvinylidene fluoride, sulfonated styrene–di-
panel) and the corresponding differential pore size distributions vinylbenzene resin, and phenol–formaldehyde resin, respectively.
calculated using 2D-NLDFT method (right panel; inset shows the Nitrogen adsorption isotherms for AC-PVDF, AC-PPy and AC-
mesopore range of PSD for AC-SDVB) for all carbons studied. AC- SDVB were reported by Choma et al. (2015)
PPy, AC-PVDF, AC-SDVB and OMC-PF denote carbons prepared

Table 1 Structural parameters of the carbons studied


Sample SDFT (m2 g-1) SBET (m2 g-1) Vt (cm3 g-1) V2 (cm3 g-1) V1 (cm3 g-1) w1 (nm) w2 (nm) Microporosity (%)

AC-PPy 1990 1810 0.87 0.78 0.62 0.6 – 90


AC-PVDF 1980 2920 1.64 1.47 0.46 0.9 2.1 90
AC-SDVB 1960 2480 1.33 1.07 0.64 0.8 1.6 80
OMC-PF 840 660 0.65 0.22 0.18 0.6 1.2 34
SDFT—specific surface area calculated using 2D-NLDFT method for carbons with slit-shaped pores under assumptions of energetic heterogeneity
and geometrical corrugation of the surface implemented in SAIEUS software; SBET—apparent specific surface area calculated by the Brunauer–
Emmett–Teller method using low-temperature nitrogen adsorption data in the relative pressure range of 0.05–0.20 and nitrogen cross-section
area of 0.162 nm2; Vt—single-point pore volume calculated by conversion of the nitrogen adsorption at a relative pressure &0.99 to the volume
of liquid nitrogen at the experiment temperature; V2—volume of micropores and small mesopores calculated by integration of PSD under the
first and second peak in the range up to the endpoint of the second peak; V1—volume of small micropores (ultramicropores) calculated by
integration of PSD under the first peak in the range up to the PSD minimum after the first peak; w1—size of small micropores (ultramicropores)
estimated as the maximum of the first PSD peak; w2—size of large micropores estimated as the maximum of the second PSD peak; Microp-
orosity-ratio of V2 to Vt expressed in %. Notation of the samples as in Fig. 1. Parameters for AC-PVDF, AC-PPy and AC-SDVB are taken from
Choma et al. (2015)

well, some with broad PSDs extending up to small meso- isotherms and Table 2 lists the CO2 uptakes for all mate-
pores. The peaks’ maxima in this range (w2) vary from 1.2 rials at each temperature. The CO2 uptakes recorded at
to 2.1 nm. The volume of all micropores and small 0 °C and 1 atm are in the range of 3.5–6.9 mmol g-1, and
mesopores (if present, labeled as V2) range from 0.22 to those at 60 °C are in the range of 1.0–1.9 mmol g-1. AC-
1.47 cm3 g-1. In addition, OMC-PF and AC-SDVB also PPy consistently scored the highest uptake at each tem-
possess some amounts of mesopores, as apparent form their perature, while the order of the other sorbents changed
PSDs. depending on the temperature. To start with, the uptakes at
Extensive CO2 adsorption studies were performed for all 0 °C follow the order: AC-PPy [ AC-SDVB [ AC-
carbons studied. CO2 adsorption measurements were done PVDF [ OMC-PF, but at higher temperatures, OMC-PF
at multiple temperatures from 0 to 60 °C and the results slowly moves from the fourth to the third place at 10 °C,
were used to calculate the temperature dependence of CO2 and then to the second position at 50 °C. The final uptake
uptake, working capacity of each adsorbent, and isosteric order, at 60 °C, ends up being AC-PPy [ OMC-PF [ AC-
heat of CO2 adsorption. Figure 2 shows all CO2 adsorption SDVB [ AC-PVDF. These interesting changes are

123
Adsorption (2016) 22:581–588 585

A 8 B 5
0oC
o
0C
o
AC-PPy 10 C 4 AC-PVDF o
10 C

CO 2 adsorption (mmol g )
CO2 adsorption (mmol g )

-1
-1

20oC o
20 C
25 C o 3 o
25 C
4 30ooC 30 C
o
40 C
2 o
50oC 40 C
o
60oC 50 C
o
2 60 C
1

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Pressure (atm) Pressure (atm)
C 7 o
0C D 4 o
0C
o
6 10 C
AC-SDVB CO2 adsorption (mmol g ) 20 C
o
OMC-PF
-1
CO2 adsorption (mmol g )

3 o
-1

o
5 10 C 25 C
o
30 C
o
4 20 C 40oC
25oC 2
30oC 50oC
3 60oC
40oC

2 50oC
o 1
60 C
1

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Pressure (atm) Pressure (atm)

Fig. 2 CO2 adsorption isotherms measured at 0–60 °C for a AC-PPy, b AC-PVDF, c AC-SDVB, and d OMC-PF. Notation of the samples as in
Fig. 1

Table 2 CO2 uptakes in


Sample CO2 uptake (mmol g-1)
0–60 °C temperature range for
the carbons studied 0 °C 10 °C 20 °C 25 °C 30 °C 40 °C 50 °C 60 °C

AC-PPy 6.92 5.69 4.64 3.92 3.70 2.98 2.32 1.89


AC-PVDF 3.79 3.09 2.62 2.30 1.95 1.46 1.22 0.96
AC-SDVB 5.78 4.19 3.34 2.99 2.65 2.23 1.64 1.28
OMC-PF 3.46 3.19 2.84 2.59 2.32 1.92 1.65 1.38
CO2 uptake at the specified temperature and pressure of 1 atm; Notation of the samples as in Fig. 1

attributed to the presence of extremely small micropores in adsorption as compared to the values obtained for other
the OMC-PF carbon that retain CO2 much stronger and carbons (see Fig. 4).
result in much slower decay of the CO2 adsorption with Figure 3 shows the CO2 uptakes at 0.0013 atm
temperature. This is supported by much higher isosteric (1 mmHg) and 1 atm (760 mmHg) as functions of tem-
heat of CO2 adsorption for OMC-PF at low values of perature for all carbons studied. The uptakes show almost

123
586 Adsorption (2016) 22:581–588

10 To complement the CO2 adsorption studies, the isosteric


1 atm
heats of adsorption were calculated based on the CO2
adsorption and the Clausius–Clapeyron equation. Figure 4
1 shows the calculated heats of adsorption as functions of the
CO2 uptake (mmol g )
-1

AC-PPy
CO2 uptake for all carbons studied. The noticeable small
AC-PVDF
AC-SDVB
steps in the plots at higher values of the CO2 uptake are due
0.1 OMC-PF to the decreasing number of the CO2 isotherms used to
AC-PPy calculate qst. These steps represent a sudden decrease in the
AC-PVDF
AC-SDVB
accuracy not the actual changes. As expected, the heats of
0.01 OMC-PF adsorption decrease with the amount of adsorbed CO2, but
the values and their decline are quite different among the
0.0013 atm carbons studied. For instance, the highest initial heat of
0.001 adsorption (32.9 kJ mol-1 at CO2 uptake of 0.1 mmol) is
observed for OMC-PF, followed by those for AC-PPy
0 10 20 30 40 50 60 70
Temperature (°C) (29.2 kJ mol-1), AC-SDVB (25.5 kJ mol-1), and AC-
PVDF (24.9 kJ mol-1). However, the fastest decreases in
Fig. 3 CO2 uptakes at 0.0013 atm (hollow symbols) and 1 atm (full qst, 35 and 29 %, in the range of 0–2.5 mmol are observed
symbols) as functions of temperature for all carbons studied. The
ordinate axis is in logarithmic scale. Notation of the samples as in
for AC-PVDF and OMC-PR, respectively, while the cor-
Fig. 1 responding decreases for AC-PPy and AC-SDVB are about
14 %.
perfect exponential dependence on temperature in the
0–60 °C interval (note the ordinate axis is in logarithmic
scale). For all straight lines in Fig. 3 the correlation coef-
ficient exceeds 0.99. The exponential dependence has been 4 Conclusions
previously reported by Ludwinowicz and Jaroniec (2015a)
and used to estimate the working capacity in a CO2 pres- Four carbon adsorbents prepared from different polymers:
sure-swing adsorption process (PSA). Aaron and Tsouris polypyrrole (AC-PPy), polyvinylidene fluoride (AC-
(2005) described PSA where adsorption occurs at 30 °C PVDF), sulfonated styrene–divinylbenzene resin (AC-
and 2 atm and desorption is carried out at 60 °C and SDVB), and phenol–formaldehyde resin (OMP-PF) were
0.0013 atm. Here, we calculate the working capacity based examined as potential adsorbents for CO2 capture. CO2
on this information, but instead of adsorption at 2, 1 atm is adsorption isotherms measured at different temperatures
used (the working capacity at 2 atm would be even higher). between 0 and 60 °C range were used to evaluate the
Table 3 lists the adsorption and desorption uptakes and the adsorption properties such as CO2 uptakes at different
final working capacity. temperatures, working capacity, and isosteric heat of
As the result of very small uptake under desorption adsorption in relation to the polymer used as carbon
conditions, the working capacities are virtually equal to the precursor.
CO2 uptakes at the adsorption conditions. Consequently, All carbons studied showed the well developed porous
AC-PPy scored the highest working capacity of structures, either exclusively microporous (AC-PPy and
3.70 mmol g-1. The other materials achieved values in the AC-PVDF) or micro-mesoporous (AC-SDVB and OMC-
range 1.95–2.65 mmol g-1, which are comparatively high PF). The highest CO2 uptake at 0 °C, 6.92 mmol g-1, was
as well (Samanta et al. 2012). obtained for AC-PPy followed by those for AC-SDVB,

Table 3 Adsorption and desorption uptakes, and the working capacity for a CO2 pressure-swing adsorption process calculated for all carbons
studied
Sample Adsorption uptake (mmol g-1) Desorption uptake (mmol g-1) Working capacity (mmol g-1)

AC-PPy 3.70 0.0018 3.70


AC-PVDF 1.95 0.0006 1.95
AC-SDVB 2.65 0.0008 2.65
OMC-PF 2.32 0.0029 2.32
Adsorption uptake denotes CO2 uptake at 30 °C and 1 atm; Desorption uptake denotes CO2 uptake at 60 °C and 0.0013 atm; the working
capacity was calculated as a difference between adsorption and desorption uptakes; Notation of the samples as in Fig. 1

123
Adsorption (2016) 22:581–588 587

A B 26

Isosteric heat of adsorption CO2 (kJ mol )


30

-1
Isosteric heat of adsorption CO2 (kJ mol )
-1

24
29 AC-PPy AC-PVDF

28 22

27 20

26 18

25
16

0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
-1 -1
CO 2 uptake (mmol g ) CO2 uptake (mmol g )
C D

Isosteric heat of adsorption CO2 (kJ mol )


26
-1
Isosteric heat of adsorption CO2 (kJ mol )
-1

32
25 AC-SDVB OMC-PF

24
28

23

22 24

21

0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
-1
CO2 uptake (mmol g ) CO2 uptake (mmol g-1)

Fig. 4 Isosteric heat of CO2 adsorption for a AC-PPy, b AC-PVDF, c AC-SDVB, and d OMC-PF calculated from CO2 isotherms adsorption
measured at 0, 10, 20, 25, 30, 40, 50 and 60 °C

AC-PVDF, and OMC-PF. At higher temperatures, AC-PPy conditions, making them viable candidates for carbon
consistently scored best CO2 uptakes among carbons capture and other adsorption and environmental-related
studied, while the order of the remaining samples changed applications.
with temperature. The temperature dependence of CO2
uptake in 0–60 °C range followed an exponential trend, as Acknowledgments J.C. acknowledges the National Science Centre
(Poland) for support of this research under grant 2013/09/B/ST5/
previously reported by Ludwinowicz and Jaroniec (2015a). 00076.
Based on these data, the working capacities for all adsor-
bents were calculated assuming a pressure-swing CO2
adsorption process with adsorption at 1 atm and 30 °C and
desorption at 0.0013 atm and 60 °C. The resulting values References
range from 1.95 to 3.70 mmol g-1, with the best result for
AC-PPy. Overall, the carbons studied were highly micro- Aaron, D., Tsouris, C.: Separation of CO2 from flue gas: a review.
porous, which resulted in high CO2 uptakes at ambient Sep. Sci. Technol. 40, 321–348 (2005)

123
588 Adsorption (2016) 22:581–588

Arienillas, A., Rubiera, F., Parra, J.B., Ania, C.O., Pils, J.J.: Surface Lee, S., Park, S.: Determination of the optimal pore size for improved
modification of low cost carbons for their application in the CO2 adsorption in activated carbon fibers. J. Colloid Interface
environmental protection. Appl. Surf. Sci. 252, 619–624 (2005) Sci. 389, 230–235 (2013)
Brunauer, S., Emmett, P.H., Teller, E.: Adsorption of gases in Liu, J., Thallapally, P.K., McGrail, B.P., Brown, D.R., Liu, J.:
multimolecular layers. J. Am. Chem. Soc. 60, 309–319 (1938) Progress in adsorption-based CO2 capture by metal-organic
Choma, J., Kalinowska, A., Jedynak, K., Jaroniec, M.: Reproducibil- frameworks. Chem. Soc. Rev. 41, 2308–2322 (2012)
ity of the synthesis and adsorption properties of ordered Liu, J., Wickramaratne, N.P., Qiao, S.Z., Jaroniec, M.: Molecular-
mesoporous carbons obtained by the soft-templating method based design and emerging applications of nanoporous carbon
(in Polish). Ochr. Srodowiska 34(3), 3–10 (2012) spheres. Nat. Mater. 14, 763–774 (2015)
Choma, J., Osuchowski, L., Jaroniec, M.: Properties and applications Ludwinowicz, J., Jaroniec, M.: Potassium salt-assisted synthesis of
of activated carbons obtained from polymeric materials: a review highly microporous carbon spheres for CO2 adsorption. Carbon
(in Polish). Ochr. Srodowiska 36(2), 3–16 (2014a) 82, 297–303 (2015a)
Choma, J., Osuchowski, L., Marszewski, M., Jaroniec, M.: Highly Ludwinowicz, J., Jaroniec, M.: Effect of activating agents on the
microporous polymer-based carbons from CO2 and H2 adsorp- development of microporosity in polymeric-based carbon for
tion. RSC Advances 4, 14795–14802 (2014b) CO2 adsorption. Carbon 94, 673–679 (2015b)
Choma, J., Stachurska, K., Osuchowski, L., Dziura, A., Jaroniec, M.: Marsh, H., Rodriguez-Reinoso, F.: Activated Carbon. Elsevier,
Carbon dioxide adsorption on activated carbons obtained from Amsterdam (2006)
polymeric precursors (in Polish). Ochr. Srodowiska 37(4), 3–8 Migahead, M.A., Abdul-Raheim, A.M., Atta, A.M., Brostow, W.:
(2015) Synthesis and evaluation of a new water soluble corrosion
De Canck, E., Ascoop, I., Sayari, A., Van Der Voort, P.: Periodic inhibitor from recycled poly(ethylene terephthalate). Mater.
mesoporous organosilicas functionalized with a wide variety of Chem. Phys. 121, 208–214 (2010)
amines for CO2 adsorption. Phys. Chem. Chem. Phys. 15, Mishra, S., Goje, A.S., Zope, V.S.: Chemical recycling, kinetics, and
9792–9799 (2013) thermodynamics of poly(ethylene terephthalate) (PET) waste
De Souza, L.K.C., Wickramaratne, N.P., Ello, A.S., Costa, M.J.F., da powder by nitric acid hydrolysis. Polym. React. Eng. 11, 79–99
Costa, C.E.F., Jaroniec, M.: Enhancement of CO2 adsorption on (2003)
phenolic resin-based mesoporous carbons by KOH activation. Myers, A.L.: Thermodynamics of adsorption in porous materials.
Carbon 65, 334–340 (2013) AIChE J. 48, 145–160 (2002)
Górka, J., Jaroniec, M.: Hierarchically porous phenolic resin-based Patel, H.A., Karadas, F., Canlier, A., Park, J., Deniz, E., Jung, Y.,
carbons obtained by block copolymer-colloidal silica templating Atilhan, M., Yavuz, C.T.: High capacity carbon dioxide
and post-synthesis activation with carbon dioxide and water adsorption by inexpensive covalent organic polymers. J. Mater.
vapor. Carbon 49, 154–160 (2011) Chem. 22, 8431–8437 (2012)
Gupta, V.K., Nayak, A., Agarwal, S., Tyagi, I.: Potential of activated Piacentini, R.D., Mujumdar, A.S.: Climate change and drying of
carbon from waste rubber tire for the adsorption of phenolics: agricultural products. Drying Technol. 27, 629–635 (2009)
effect of pre-treatment conditions. J. Colloid Interface Sci. 417, Samanta, A., Zhao, A., Shimizu, G.K.H., Sarkar, P., Gupta, R.: Post-
420–430 (2014) combustion CO2 capture using solid sorbents: a review. Ind. Eng.
Heydari-Gorji, A., Belmabkhout, Y., Sayari, A.: Polyethyleneimine- Chem. Res. 51, 1438–1463 (2012)
impregneted mesoporous silica: effect of amine loading and Seredych, M., Jagiello, J., Bandosz, T.J.: Complexity of CO2
surface alkyl chains on CO2 adsorption. Langmuir 27, adsorption on nanoporous sulfur-doped carbons—is surface
12411–12416 (2011) chemistry an important factor? Carbon 74, 207–217 (2014)
ICPP, Climate Change 2013: The Physical Science Basis. Contribu- Su, F., Lu, C.: CO2 capture from gas stream by zeolite 13x using a
tion of Working Group I to the Fifth Assessment Report of the dual-column temperature/vacuum swing adsorption. Energy
Intergovernmental Panel of Climate Change, Cambridge Univer- Environ. Sci. 5, 9021–9027 (2012)
sity Press, Cambridge, United Kingdom and New York, 2013 Thommes, M., Kaneko, K., Neimark, A.V., Olivier, J.P., Rodriguez-
Jagiello, J., Olivier, J.P.: 2D-NLDFT Adsorption models for carbon Reinoso, F., Rouquerol, J., Sing, K.S.W.: Physisorption of gases,
slit-shaped pores with surface energetical heterogeneity and with special reference to the evaluation of surface area and pore
geometrical corrugation. Carbon 55, 70–80 (2013a) size distribution (IUPAC Technical Report). Pure Appl. Chem.
Jagiello, J., Olivier, J.P.: Carbon slit pore model incorporating surface (2015). doi:10.1515/pac-2014-1117
energetical heterogeneity and geometrical corrugation. Adsorp- Wang, X., Liang, C.D., Dai, S.: Facile synthesis of ordered
tion 19, 777–783 (2013b) mesoporous carbons with high thermal stability by self-assembly
Jankowska, H., Swiatkowski, A., Choma, J.: Active Carbon. Ellis of resorcinol–formaldehyde and block copolymers under highly
Horwood Ltd., Chichester (1991) acidic conditions. Langmiur 24, 7500–7505 (2008)
Kruk, M., Jaroniec, M.: Gas adsorption characterization of ordered Wickramaratne, N.P., Jaroniec, M.: Activated carbon spheres for CO2
organic–inorganic nanocomposite materials. Chem. Mater. 13, adsorption. ACS Appl. Mater. Interfaces 5, 1849–1855 (2013)
3169–3183 (2001)

123

You might also like