You are on page 1of 15

Research Article

Adsorption Science & Technology


Surface characteristics of 2018, Vol. 36(1–2) 478–492
! The Author(s) 2017
KOH-treated commercial DOI: 10.1177/0263617417704527
journals.sagepub.com/home/adt

carbons applied for CO2


adsorption
Zofia Lendzion-Bieluń, Ł Czekajło, Daniel Sibera,
Dariusz Moszyński, Joanna Sreńscek-Nazzal,
Antoni W Morawski, Rafal J Wrobel,
Beata Michalkiewicz, Walerian Arabczyk and
Urszula Narkiewicz
West Pomeranian University of Technology, Poland

Abstract
The effect of an alkali treatment (potassium hydroxide) on the properties of a commercial
activated carbon has been studied. The aim of the treatment was to improve the adsorption
properties of the material toward carbon dioxide. In the result of the treatment, silica contained
in the raw carbon was removed and the density of the material increased. The changes in the
surface chemistry were observed as well. The treatment of the activated carbon with KOH
resulted in a complete removal of carboxy and lactone groups and a decrease of the general
content of the acidic groups (more significant than that of basic groups). Simultaneously, the
surface concentration of hydroxyl groups increased. The alkali treatment of activated carbon
resulted in an increase of carbon dioxide uptake of 14% (measured using a volumetric method
at 0 C). The adsorption of carbon dioxide on activated carbon has a mixed (physicochemical)
character and that two types of adsorption sites are present at the surface. The adsorption
energy varies roughly from 25 to 60 kJ/mol.

Keywords
Activated carbon, carbon dioxide, adsorption, surface chemistry, alkali treatment

Submission date: 21 November 2016; Acceptance date: 8 March 2017

Corresponding author:
Urszula Narkiewicz, West Pomeranian University of Technology, w Szczecinie Pulaskiego, 10 Szczecin 70-322, Poland.
Email: urszula.narkiewicz@gmail.com

Creative Commons CC-BY: This article is distributed under the terms of the Creative Commons
Attribution 4.0 License (http://www.creativecommons.org/licenses/by/4.0/) which permits any use,
reproduction and distribution of the work without further permission provided the original work is attributed as specified
on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
Lendzion-Bieluń et al. 479

Introduction
A decrease of carbon dioxide emission is one of the most important challenges today. It is
believed that it would be a remedy to inhibit the global warming of our planet and dangerous
climate changes. Fossil fuel power plants are the main source of anthropogenic CO2
emissions, but industrial production plants (e.g. chemical industry, cement industry, iron
and steel production) are also large contributors. There are a lot of solutions proposed for
carbon dioxide storage, conversion, and utilization (Song, 2006), but in all the cases the first
step has to be a CO2 capture, its separation from exhaust gases. CO2 can be separated
directly from the flue gas (so-called PCCC—postcombustion carbon capture) by
absorption, adsorption, membrane separation, cryogenic distillation, and solidification
(Fu and Gundersen, 2012).
The existing processes used for PCCC and based mainly on absorption in liquids
(as amines, potassium carbonate, or methanol) are costly, inefficient, and may have
inherent environmental problems. There is a need to develop new PCCC technologies,
more cost and energy efficient and more friendly for environment. Solid sorbents can be
used in such processes, instead of traditional liquids. In a review paper on solid sorbents for
PCCC (Choi et al., 2009), such materials as zeolites, activated carbons, calcium oxides,
hydrotalcites, organic–inorganic hybrids, and metal–organic frameworks are mentioned.
In another recent review paper on PCCC using solid sorbents, the criteria for choosing
the best CO2 sorbents are determined (Samanta et al., 2012) as: adsorption capacity and
selectivity for CO2, adsorption/desorption kinetics (fast enough under operating conditions),
mechanical strength of sorbent particles, chemical stability/tolerance to impurities, low heat
needed for a regeneration, and low cost of the adsorbent ($15/kg sorbent—would be not
economical).
As carbonaceous materials can fulfill most of the above criteria, they seem to be a good
candidate for the application at industrial scale.
Activated carbons are well-known sorbent materials for various applications, as removal
of organic and inorganic impurities from water (Mohan and Pittman, 2006; Otowa et al.,
1997) and gases, then it is not surprising that they can also be applied for CO2 separation.
A range of raw materials can be used to obtain activated carbons, for example coals,
industrial by-products, or biomass products (as plum stones, coconut shells, or algae)
(Aravindhan et al., 2009). Characteristic features of activated carbons are high specific
surface area, well developed micropores structure, and interesting surface chemistry
created by the presence of various heteroatoms and functional groups. The surface
chemistry of the activated carbons can be modified to change their adsorption behavior.
The modification can be thermal or chemical and in the latter case it can be carried out in gas
or liquid phase. For the modification in liquid phase often nitric or sulfuric acid is applied
(or their mixture) or hydrogen peroxide. Activated carbon can be also modified during the
production process through chemical or physical activation. In the case of chemical
activation, the raw material is impregnated with a chemical compound (as KOH, K2CO3,
NaOH, Na2CO3, salts of alkali earth metals—AlCl3 and ZnCl2, and some acids as H3PO4 i
H2SO4) and then heated under inert gas.
Maroto-Valer et al. (2005) investigated the CO2 capture behavior of steam-activated
anthracite and concluded that CO2 uptake did not depend proportionally on the specific
surface area of the material. The highest CO2 adsorption capacity (65.7 mg CO2/g of
adsorbent) was reached on the sample with a specific surface area of 540 m2/g only,
480 Adsorption Science & Technology 36(1–2)

when the sample having almost doubled highest specific surface area of 1071 m2/g adsorbed
much less of CO2–40 mg CO2/g. It means that despite the specific surface area there are other
factors influencing the adsorption capacity toward CO mg CO2/g as porous structure and
surface chemistry.
Pevida et al. (2008) proved that nitrogen functionalities could improve the adsorption
capacity of activated anthracites toward carbon dioxide.
Another way to introduce a basic surface functionality, which promotes CO2 capture
capacity, can be activation with potassium hydroxide. KOH has been reported as a
carbon activator to improve the adsorption properties toward hydrogen sulfide
(Sitthikhankaew et al., 2014) and nitrogen oxides (Lee et al., 2002) but also for carbon
dioxide adsorption.
According to Lee et al. (2014), an activation of ordered nanoporous carbons with KOH
caused an increase of specific surface area and total pore volume, which resulted in the
enhancement of CO2 adsorption capacity. A similar conclusion was drawn by de Andres
et al. (2013); however, better results were reached using solid NaOH.
Wei et al. (2012) reported a high surface area and CO2 uptake at high pressure (3.57 MPa)
of activated carbon produced from Finger Citron residue and treated with KOH.
A biological material (olive stones (Ubago-Perez et al., 2006)) was also used for
fabrication of different granular and powder-activated carbons obtained by KOH activation.
To prepare efficient carbon dioxide sorbent with a basic surface, the doping with nitrogen
is also recommended (Xing et al., 2012). The authors showed that introduction of nitrogen
into a carbon surface facilitates the hydrogen-bonding interaction between the carbon
surface and CO2 molecules. Recently, an interesting paper of Zhou et al. (2015) was
published, reporting a superior CO2 sorption on N-doped microporous carbons derived
from direct carbonization of Kþ exchanged meta-aminophenol–formaldehyde resin. The
authors concluded that a good performance of such a sorbent toward carbon dioxide is
connected with an appropriate porous structure (uniform ultramicropores around 0.5 nm)
and N-doping. The same research group of Zhou et al. (2016) followed the research on
ultramicroporous sorbents for carbon dioxide enhanced adsorption and proposed a direct
pyrolysis of alkali salts of carboxylic phenolic resins. The size of pores in the final product
can be finely tuned depending on the applied alkali metal.
In our studies, a commercial activated carbon was modified with KOH. In the previous
paper (Sreńscek-Nazzal et al., 2015), we showed the promising porous structure of such a
commercial activated carbon, the total pore volume Vp was equal to 0.5 cm3/g with 80% of
pores occupied by micropores (Vmic ¼ 0.4 m3/g). The specific surface area (Brunauer–
Emmett–Teller (BET)) of the carbon was high and equal to 1186 m2/g. The CO2
adsorption isotherms measured at 40 C up to 40 bar showed a significant increase of the
carbon dioxide uptake caused by the addition of KOH, 1.4 times at 40 bar.
In the present paper, we are describing the properties of the same material (commercial
WG12 carbon), but activated with KOH in a microwave-assisted hydrothermal reactor,
because recently there are reports about a positive effect of microwave treatment (Kim
et al., 2014; Schwenke et al., 2015; Zhang et al., 2013). The issue of the use of
hydrothermal processes in carbon activation (HTC-based carbons) has been thoroughly
treated in the book of Titirici (2013). Both the surface area and pore volume of
hydrothermally produced carbons activated with KOH are formed mainly by micropores.
The aim of the present paper is to study the effect of KOH treatment on the surface
properties of a commercial carbon activated with KOH.
Lendzion-Bieluń et al. 481

Experimental
A sample of 3 g of commercial activated carbon WG12 (Gryfskand) was suffused with a
solution of KOH (15 g in 50 ml of water) and put into a microwave-assisted hydrothermal
reactor Magnum II (Ertec Poland). The process was carried out at 3 MPa for 15 min.
The obtained material was washed with distilled water until a neutral pH was reached
and next heated at 100 C for 24 h.

Characterization
The studies of the adsorption of carbon dioxide on the prepared carbon sample were carried
out using a temperature-programmed desorption method (TPD-CO2) with a micrometrics
equipment. The studies were performed under atmospheric pressure at three different
temperatures: 30, 0, and 20 C. Before the CO2 adsorption studies the carbon samples
were heated under helium at 300 C for 2 h. Then, the samples were cooled to the
adsorption temperature under helium. To reach a selected adsorption temperature (30, 0,
or 20 C) a liquid nitrogen cooling system was applied (Cryocooler). The CO2 adsorption was
performed with a stream of CO2 passing through the adsorbent bed with a rate of 50 cm3/min
for 30 min. After the adsorption CO2 was replaced by helium, passing through the bed for
30 min at the adsorption temperature, to remove carbon dioxide not bounded to the surface.
The TPD-CO2 process was conducted by applying an increase of temperature with a rate
of 15, 20, and 25 /min under a flux of helium through the adsorbent bed. Three rates of
temperature increase were applied to determine a desorption energy (Ed) of CO2 from the
carbon surface.
To determine the energy of CO2 desorption the first raw kinetics was applied according to
equation (1)
lnðb=T2max Þ ¼ Ed =RTmax þ lnðEd =ðR: CÞÞ ð1Þ

in which:
b—rate of the temperature increase ( /min),
Ed—desorption energy (J/mol K),
R—gas constant,
Tp—a temperature, in which the desorption rate is the highest (K)

The energies of desorption were determined as tangents to the lines traced according
to the dependence ln (b/T2max) ¼ f(1/Tmax) for all cases of adsorption on carbon under
the study.
The textural properties of the ACs were determined by physical adsorption of N2 at 77 K
and CO2 at 273 K using a Quadrasorb apparatus (Quantachrome Instruments). Before the
experiments, the samples were outgassed under vacuum at 260 C overnight. The specific
surface area was measured by the multipoint BET method.
The helium density of the prepared samples was measured under helium using a Micro-
Ultrapyc 1200e equipment at the pressure of 17 psi. The samples were purged with helium for
20 min.
The presence of functional groups was assessed using a modified Boehm (1994, 2002)
titration method. The modification of the method involved an adaptation of the analysis
procedure to the small amount of the samples.
482 Adsorption Science & Technology 36(1–2)

A determination of oxygen surface groups using a method of alkalimetric titration is


based on the assumption that acid constants of the carboxy, lactone, and phenyl groups
differ greatly from one another (the difference should be of several orders of magnitude).
In such a case it is possible to titrate the different groups by their neutralization with
appropriate bases.
To carry out the titration the samples of 0.2 g were taken and suffused with the solutions
of NaHCO3, Na2CO3, KOH, and HCl. In the next step of the procedure, the suspensions
were mixed in the closed small flasks for 48 h. Next, the samples remained for 24 h for a
decantation. Finally, the samples were filtered through a filter paper and the solutions were
titrated using methyl orange and phenolphthalein.
The FTIR spectrum of the tested carbons was recorded with the Thermo Scientific Nicolet
380 spectrometer.
The surface composition was studied by X-ray photoelectron spectroscopy (XPS). The X-ray
photoelectron spectra were obtained using Al Ka (hn ¼ 1486.6 eV) radiation with a Prevac
system equipped with Scienta SES 2002 electron energy analyzer operating at constant
transmission energy. Survey spectra were collected with pass energy of 100 eV, and high-
resolution spectra were acquired with pass energy of 50 eV. The spectrometer was calibrated
by using the following photoemission lines (with reference to the Fermi level): EB Cu 2p3/
2 ¼ 932.8 eV, EB Ag 3d5/2 ¼ 368.3 eV, and EB Au 4f7/2 ¼ 84.0 eV. The instrumental resolution,
as evaluated by the full width at half maximum of the Ag 3d5/2 peak, was 1.0 eV. The samples
were loosely placed into a grooved molybdenum sample holder. The analysis chamber during
experiments was evacuated to better than 51010 mbar. The surface composition of the samples
was obtained on the basis of the peak area intensities using the sensitivity factor approach and
assuming homogeneous composition of the surface layer. Obtained spectra were processed with
use of CasaXPS software. Binding energy corrections, background subtraction, and
deconvolution of the spectra have been applied. The phase composition of the samples was
determined by X-ray Diffraction (XRD) method with the PANalytical Empyrean X-ray
diffractometer using a Cu Ka radiation (a ¼ 1.5418Å) at room temperature.

Results and discussion


The comparison of the effect of the XRD spectra of the pristine and KOH-modified WG12
carbon is shown in Figure 1. There are no significant differences between the two spectra.
No phases containing any potassium compounds can be detected using XRD method for the
carbon treated with KOH.
The results of the Boehm’s titration are shown in Table 1. The treatment of the activated
carbon with KOH resulted in a complete removal of carboxy and lactone groups. A decrease
of the general content of the acidic groups was at the level of 9.35% roughly and a decrease
of the general content of basic groups was 8.01%.
The effect of the KOH treatment on the IR spectra is shown in Figure 2.
Comparing the IR spectrum of commercial coal WG12 and WG12-KOH15 after
chemical modification we do not notice significant differences. In both spectra in the
range of 3000–3750 cm1 wide band derived from the group –OH is visible. The intensity
of this band is greater after modification with KOH.
The peaks in the range 2700–3000, 1250–1500, and 500–700 cm1 can be attributed to the
C–H bending vibration and C–O stretching. The adsorption peak at 1630 cm1 corresponds
to the band C¼C.
Lendzion-Bieluń et al. 483

Figure 1. The effect of the KOH on the XRD spectra of the WG12 carbon.
XRD: X-ray Diffraction.

Table 1. An average content of the functional groups (mmol/g).

WG-12
Reactant Group WG-12 15gKOH

Distilled H2O Active acidity 0.0000 0.0000


0.05 M NaHCO3 The strongest carboxy groups, 0.0000 0.0000
e.g. –COOH
0.05 M Na2CO3 Carboxy and lactone groups 0.5703 0.0000
0.05 M NaOH General content of acidic groups 0.1155 0.1047
0.05 M HCl General content of basic groups 0.3893 0.3581

The treatment of the carbon with potassium hydroxide caused a decrease of the helium
density, resulting from some mineral impurities (ash) removal.
The chemical composition of the surface was studied by XPS. In Figure 4 survey
XPS spectra acquired before and after KOH treatment are shown. On the surface of the
sample before treatment, carbon, oxygen, and silicon atoms were identified, while after KOH
treatment the XPS signal originating from silicon atoms disappeared. Atomic concentrations
of carbon, oxygen, and silicon present on the surface of carbon WG12 are shown in Table 2.
High-resolution XPS spectra give additional insight into the chemical state of the surface
of the WG12 samples. The maximum of XPS Si 2p line is at 104.0 eV the position typical for
silica (Chourasia, 2006). In Figure 5 the XPS spectrum of O 1s transition is shown.
484 Adsorption Science & Technology 36(1–2)

Figure 2. An influence of the modification of carbon with KOH on the IR spectra.

Figure 3. Decrease of the helium density with a KOH addition.

The deconvolution of the O 1s line acquired from the surface of carbon WG12 before KOH
treatment is shown as shaded areas. The component of highest intensity is located at the
binding energy of 533.6 eV. This position is characteristic for oxygen atoms in silicon oxide
SiO2 (Chourasia, 2006) and confirms the presence of silica on the surface of this sample.
Lendzion-Bieluń et al. 485

C 1s

4
Intensity, MCPS

O 1s
2
before KOH treatment

1
after KOH treatment
Si 2s Si 2p
0
800 700 600 500 400 300 200 100 0
Binding energy, eV

Figure 4. Survey XPS spectra of the carbon WG12 before and after KOH treatment.

Table 2. Atomic concentration of carbon, oxygen, and silicon atoms on the


surface of carbon WG12.

C O Si
Sample at. %

WG12 90.1 6.9 2.9


WG12-KOH15 95.7 4.3 –

After KOH treatment XPS signal of silicon disappears. The XPS O 1s spectrum collected
after KOH treatment is apparently shifted to the low energy side (dashed line in Figure 5).
The deconvolution of this spectrum is shown as thin lines. It is noteworthy that the
component previously ascribed to the Si–O bonds, located at 533.6 eV, disappeared
completely. The other components corresponding to the C–O bonds (at 532.5 eV) and
C¼O bonds (at 531.4 eV) are slightly shifted within model constraints. However, the ratio
of these two components area is almost identical before and after KOH treatment. It is
concluded that KOH treatment removes SiO2 from the surface of carbon WG12.
Simultaneously, the exposure of the surface of carbon WG12 to the KOH solution does
not result in the formation of any new form of carbon–oxygen bonds. Before and after KOH
treatment only hydroxyl (CO) and carbonyl (C¼O) groups are detected on the surface.
The latter statement is supported by the analysis of high-resolution XPS C 1s spectra.
A spectrum originating from carbon WG12 before KOH treatment is shown in Figure 6.
486 Adsorption Science & Technology 36(1–2)

Figure 5. High-resolution XPS spectra of O 1s line before (thick solid line) and after (dashed line) KOH
treatment. The components of O 1s line for the former case are depicted as shaded areas while for the latter
case as thin solid lines.

Exp. data
CC/CH
200 C*-CO
C-O
C=O
COOR
satellite
150 Envelope
Intensity, kCPS

100

50

0
296 294 292 290 288 286 284 282
Binding energy, eV

Figure 6. High-resolution spectrum of XPS C 1s line from KOH-treated carbon WG12.


Lendzion-Bieluń et al. 487

The spectrum obtained from carbon WG12 after KOH treatment (not shown) is virtually
identical. The envelope of the C 1s line is typical for the carbon black (Gardner et al., 1995)
or other graphite-like specimens (Pelech et al., 2012). At the low-energy side there is a slim
and intense peak with maximum at 284.4 eV which is accompanied by a long tail extending
to the binding energy of about 293 eV. It was numerically deconvoluted into six components
shown as thin lines below the envelope of XPS data. Considering components from the
lowest binding energy one can attribute the given peak to the following chemical bindings
of carbon atoms: position 284.4 eV (black line) corresponds essentially to carbon atoms
located in graphitic rings; position 285.5 eV (brown line) is attributed to sp3 carbon
atoms, bonded either with second carbon or with hydrogen atoms; position 286.4 eV
(green line) is ascribed to a group of differently bonded carbon atoms linked to one atom
of oxygen, i.e. the functional groups as C–O–C or C–OH; position 288.0 (pink line)
corresponds to functional groups as C¼O or O–C–O; position 289.8 eV (gray line)
attributed to carbon atoms indicated by asterisk in the functional groups like C–O–C*¼O
or HO–C*¼O; position 291.4 eV (blue line) attributed to shake-up structure caused by the
p! p* transition in the graphite rings (Dieckhoff et al., 1995). The binding energy
assignments described above are based on the energy shifts given in Appendix E of the
book of Briggs and Grant (2003).
The treatment of the commercial activated carbon with potassium hydroxide did
not result in significant changes in porosity and specific surface area of the material.
In the case of the untreated commercial carbon, the total pore volume Vp was equal to
0.50 cm3/g with 80% of pores occupied by micropores (Vmic ¼ 0.40 m3/g) and the specific

Figure 7. Effect of KOH treatment on the porous structure.


488 Adsorption Science & Technology 36(1–2)

surface area was equal to 1128 m2/g. After the alkali treatment the specific surface area
increased slightly to 1181 m2/g and the total pore volume and micropore volume to 0.56
and 0.42 m3/g, respectively.
The effect of KOH treatment on the porous structure of material is shown in Figure 7.
The changes in porosity after KOH treatment are not very significant. The pores between 3
and 4.5 nm disappeared but also those between 1.3 and 2 nm. On the contrary, there are
more pores between 2 and 3 nm.
The addition of KOH resulted in an increase of carbon dioxide uptake measured using a
volumetric method at 0 C of 14% (from 3.16 to 3.67 mmol/g). With an increasing
temperature the CO2 uptake of the modified carbon dropped to 2.43 mmol/g at 25 C.
The results of TPD-CO2 measurements are shown in Figure 8. The lines of desorption
of CO2 from the surface of WG12 and from the surface of the carbon modified with a
solution of potassium hydroxide WG12-KOH15 correspond to the various rates of
temperature increase.
For the sample of unmodified carbon two desorption peaks are visible—one with a
maximum at 18.6 C (temperature increase rate 25 /min), and a second one with a
maximum at 119.9 C. The modification of carbon with potassium hydroxide caused a shift
of the first maximum to higher temperatures (peak at 27.2 C, increase rate 25 /min), whereas
the second peak disappeared.
An increase of the CO2 adsorption temperature to 0 C (Figure 9) caused a shift of the first
maximum of the desorption peak to temperature of 60 C. The second maximum in the case
of unmodified carbon appears at about 115 C, whereas in the case of KOH-modified
carbon—at about 195 C.
The shift of maxima (at the same rate of temperature increase) into the direction of higher
temperatures is connected with an increase of binding energy of carbon dioxide molecules to
the surface.
The desorption characteristics after an adsorption of carbon dioxide at 20 C (Figure 10)
is similar to those at lower adsorption temperatures. The maxima of desorption peaks were
shifted to higher temperatures, as it was in the previous case.

Figure 8. TPD-CO2 after the adsorption of carbon dioxide at 30 C on (a) carbon WG12 and (b) carbon
modified with KOH WG12-KOH15.
Lendzion-Bieluń et al. 489

Figure 9. TPD-CO2 after the adsorption of carbon dioxide at 0 C on (a) carbon WG12 and (b) carbon
modified with KOH WG12-KOH15.

Figure 10. TPD-CO2 after the adsorption of carbon dioxide at 20 C on (a) carbon WG12 and (b) carbon
modified with KOH WG12-KOH15.

On the basis of the performed experiments the desorption energy of carbon dioxide from
the surface was determined as well as the volume of CO2 (Table 3). The energy E1 was
determined on the basis of the desorption peak at lower temperature, E2 corresponds to a
higher temperature. At the temperature of 30 C the adsorption energy is equal to 25.8 and
26.9 kJ/mol on the unmodified WG12 i and modified WG12þKOH15, respectively. Taking
into account the value of the adsorption energy, it can be stated that the adsorption of
carbon dioxide on activated carbon has a physical character (physisorption). However, the
value of adsorption energy 73.8 kJ/mol can be considered as a chemisorption. Then, it can be
concluded that the adsorption of carbon dioxide on activated carbon has a mixed (physical/
chemical) character and that two types of adsorption sites are present at the surface. It can
also be concluded that on the unmodified carbon there are adsorption sites with higher bond
strength. Adsorption at 0 and 20 C occurs at adsorption sites with a bond strength of 46–58
490 Adsorption Science & Technology 36(1–2)

Table 3. The effect of KOH treatment on energy of carbon dioxide adsorption.

WG12 WG12-KOH15

Temperature E1 E2 VCO2 E1 E2 VCO2


adsorption ( C) (kJ/mol) (kJ/mol) (cm3/g) (kJ/mol) (kJ/mol) (cm3/g)

30 25.8 73.8 0.878 26.9 – 3.030


0 46.5 53.9 0.272 45.4 57.0 0.405
20 58.6 75.5 0.240 58.4 64.2 0.216

and 54–75 kJ/mol, respectively. Modification of carbon with KOH resulted in higher CO2
uptake at 30 and 0 C. At 20 C the adsorbed volume is similar for both modified and
unmodified carbon.

Conclusions
The treatment of commercial carbon WG12 with KOH in the microwave-assisted
hydrothermal reactor results in increased carbon dioxide uptake, even if the modified
material was completely free of KOH (proved using XPS and XRD). According to the
XPS studies, the elimination of silica from the surface of carbon WG12 occurred as well
due to the KOH treatment.
Only hydroxyl (C–O) and carbonyl (C¼O) groups are detected on the surface before and
after KOH treatment. Only hydroxyl (C–O) and carbonyl (C¼O) groups are detected on the
surface before and after KOH treatment; however, the treatment does not affect the chemical
state of oxygen atoms bound to the carbon.
The treatment with KOH is connected with an increased concentration of hydroxyl
groups on the surface, demonstrated using IR spectroscopy.
According to Boehm’s analysis, the treatment of the activated carbon with KOH resulted
in a complete removal of carboxy and lactone groups. A decrease of the general content of
the acidic groups was at the level of 9.35% roughly, and a decrease of the general content of
basic groups was 8.01%.
The adsorption of carbon dioxide on the commercial activated carbon can have both chemical
and physical character and both types of adsorption sites are present at the carbon surface.
The general conclusion is that a treatment of activated commercial carbon with potassium
hydroxide causes an increase of hydroxyl groups on the surface, resulting in an increased
carbon dioxide uptake.

Declaration of Conflicting Interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or
publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/or
publication of this article: The studies presented in the paper were performed within the framework of
Lendzion-Bieluń et al. 491

Project Contract No. Pol-Nor/237761/98/2014 funded from the Polish-Norwegian Research


Programme operated by the National Centre for Research and Development under the Norwegian
Financial Mechanism 2009–2014.

References
de Andres JM, Orjales L, Narros A, et al. (2013) Carbon dioxide adsorption in chemically
activated carbon from sewage sludge. Journal of the Air and Waste Management Association
63(5): 557–564.
Aravindhan R, Raghava Rao J and Unni Nair B (2009) Preparation and characterization of activated
carbon from marine macro-algal biomass. Journal of Hazardous Materials 162: 688–694.
Boehm HP (1994) Some aspects of the surface chemistry. Carbon 32: 759–769.
Boehm HP (2002) Surface oxides on carbon and their analysis: A critical assessment. Carbon 40:
145–149.
Briggs D and Grant JT (2003) Surface Analysis by Auger and X-ray Photoelectron Spectroscopy.
Charlton and Manchester: IM Publications and SurfaceSpectra Limited.
Choi S, Drese JH and Jones CW (2009) Adsorbent materials for carbon dioxide capture from large
anthropogenic point sources. Chemistry and Sustainability, Energy and Materials 2: 796–854.
Chourasia AR (2006) Core level XPS spectra of silicon dioxide using zirconium and magnesium
radiations. Surface Science Spectra 13: 48–57.
Dieckhoff S, Schlett V, Possart W, et al. (1995) XPS studies on thin polycyanurate films on silicon
wafers. Fresenius’ Journal of Analytical Chemistry 353: 278–281.
Fu C and Gundersen T (2012) Carbon capture and storage in the power industry: Challenges and
opportunities. Energy Procedia 16: 1806–1812.
Gardner SD, Singamsetty CSK, Booth GL, et al. (1995) Surface characterization of carbon-fibers using
angle-resolved XPS and ISS. Carbon 33: 587–595.
Kim T, Lee J and Lee K-H (2014) Microwave heating of carbon-based solid materials. Carbon Letters
15: 15–24.
Lee S-Yi, Yoo HM, Rhee KY, et al. (2014) Synthesis, characterization, and KOH activation of
nanoporous carbon for increasing CO2 adsorption capacity. Research on Chemical Intermediates
40(7): 2535–2542.
Lee Y-W, Choi D-K and Park J-W (2002) Characteristics of NOx adsorption and surface chemistry on
impregnated activated carbon. Separation Science and Technology 37(4): 937–956.
Maroto-Valer MM, Tang Z and Zhang Y (2005) CO2 capture by activated and impregnated
anthracites. Fuel Processing Technology 86(14–15): 1487–1502.
Mohan D and Pittman CU (2006) Activated carbons and low cost adsorbents for remediation of
tri- and hexavalent chromium from water. Journal of Hazardous Materials 137: 762–811.
Otowa T, Nojima Y and Miyazaki T (1997) Development of KOH activated high surface area carbon
and its application to drinking water purification. Carbon 35(9): 1315–1339.
Pelech I, Narkiewicz U, Moszyński D, et al. (2012) Simultaneous purification and functionalization of
carbon nanotubes using chlorination. Journal of Materials Research 27: 2368–2374.
Pevida C, Plaza MG, Arias B, et al. (2008) Surface modification of activated carbons for CO2 capture.
Applied Surface Science 254(22): 7165–7172.
Samanta A, Zhao A, Shimizu GKH, et al. (2012) Post-combustion CO2 capture using solid sorbents:
A review. Industrial and Engineering Chemistry Research 51: 1438–1463.
Schwenke AM, Hoeppener S and Schubert US (2015) Synthesis and modification of carbon
nanomaterials utilizing microwave heating. Advanced Materials 27(28): 4113–41.
Sitthikhankaew R, Chadwick D, Assabumrungrat S, et al. (2014) Effect of KI and KOH impregnations
over activated carbon on H2S adsorption performance at low and high temperatures. Separation
Science and Technology 49(3): 354–366.
492 Adsorption Science & Technology 36(1–2)

Song C (2006) Global challenges and strategies for control, conversion and utilization of CO2 for
sustainable development involving energy, catalysis, adsorption and chemical processing. Catalysis
Today 115(2): 2–32.
Sreńscek Nazzal J, Glonek K, Mlodzik J, et al. (2015) Increase the microporosity and CO2 adsorption
of a commercial activated carbon. Applied Mechanics and Materials 749: 17–21.
Titirici MM (2013) Sustainable Carbon Materials from Hydrothermal Processes. London: John Wiley &
Sons.
Ubago-Perez R, Carrasco-Marin F, Fairen-Jimenez D, et al. (2006) Granular and monolithic activated
carbons from KOH-activation of olive stones. Microporous and Mesoporous Materials 92: 64–70.
Wei D, Yuchen L, Wei S, et al. (2012) Preparation and CO2 sorption of a high surface area activated
carbon obtained from the KOH activation of finger citron residue. Adsorption Science and
Technology 30(2): 183–191.
Xing W, Liu C, Zhou Z, et al. (2012) Superior CO2 uptake of N-doped activated carbon through
hydrogen-bonding interaction. Energy and Environmental Science 5: 7323–7327.
Zhang L, Mi M, Li B, et al. (2013) Modification of activated carbon by means of microwave heating
and its effects on the pore texture and surface chemistry. Research Journal of Applied Sciences,
Engineering and Technology 5(5): 1836–1840.
Zhou J, Li Z, Xing W, et al. (2015) N-doped microporous carbons derived from direct carbonization of
Kþ exchanged meta-aminophenol-formaldehyde resin for superior CO2 sorption. Chemical
Communications 51: 4591–4594.
Zhou J, Li Z, Xing W, et al. (2016) A new approach to tuning carbon ultramicropore size at sub-
angstrom level for maximizing specific capacitance and CO2 uptake. Advanced Functional Materials
26(44): 7955–7964.

You might also like