Você está na página 1de 8

Fuel Processing Technology 91 (2010) 17681775

Contents lists available at ScienceDirect

Fuel Processing Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / f u p r o c

Oxidation of activated carbon by dry and wet methods


Surface chemistry and textural modications
Josefa Jaramillo a, Pedro Modesto lvarez a, Vicente Gmez-Serrano b,
a
b

Dpto. de Ingeniera Qumica y Qumica Fsica, Universidad de Extremadura, Badajoz 06071, Spain
Dpto. de Qumica Orgnica e Inorgnica, Universidad de Extremadura, Badajoz 06071, Spain

a r t i c l e

i n f o

Article history:
Received 27 March 2010
Received in revised form 22 July 2010
Accepted 24 July 2010
Keywords:
Activated carbon
Dry and wet oxidations
Surface chemistry modication
Textural changes

a b s t r a c t
The effects of dry and wet oxidation treatments of activated carbon (AC) on the surface chemistry and
porous structure are studied. Using cherry stones (CS), AC was rst prepared by carbonization at 900 C for
2 h in N2 and activation at 850 C for 2 h in CO2. Then, the resulting AC was oxidized in O2(air) or O3
atmosphere and with HNO3 and H2O2 solutions. The acidicbasic surface sites were analyzed by FT-IR
spectroscopy, Boehm method, and pH of the point of zero charge (pHpzc) and the porous structure by N2
adsorption and mercury porosimetry. It has been found that the oxidizing agent, under specic reaction
conditions, rather than whether it was a gas or a solute in aqueous solution, is the main factor that controls
the changes produced in the surface chemistry and porous structure of AC. O3 and HNO3 are the most
effective oxidants to form acidic oxygen surface groups. However, the content of basic groups decreases for
the four oxidants, the effect being much stronger for HNO3. A microporosity reduction is also observed,
which is more important for O2(air) and especially for HNO3 than for O3 and H2O2. The percentage of
microporosity loss is as high as 43.3 for HNO3. Mesoporosity signicantly develops, whereas macroporosity
usually remains practically unchanged. Dry oxidation of AC at 100 C in O3 has proved to be the most
promising method to increase the content of acidic oxygen surface groups in the material without greatly
decreasing the content of basic sites and microporosity and with a signicant mesoporosity development.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Activated carbon is a carbon porous material with a large number
of present-day and potential industrial applications related to the
energy and fuel storage and use. Such applications include the use in
cells [18], supercapacitors [912], batteries [13,14] and canisters
[15], fuel storage [1622], production [2325] and beneciation [2631],
gas separation [32,33], and environmental protection [3437]. Since the
performance of AC in most applications is inuenced by its surface
chemistry [22,28,3844], the modication of this AC property has
been frequently the target of a variety of AC beneciation treatments.
Oxidation is one of the most conventional modications used for ACs.
It is mainly used to introduce carbonoxygen surface groups in AC.
Oxidation methods involve the utilization of oxidizing gases (i.e. oxygen, steam, carbon dioxide, etc.) or oxidizing solutions (i.e., nitric acid,
hydrogen peroxide, chlorine water, etc.) [45]. However, the oxidation
of AC from the gas phase or liquid phase may produce changes in
the porous structure of the material which may also affect its behavior.
As a result, research aiming at modifying the surface chemistry of
activated carbon must be coupled with a parallel study of the porous

Corresponding author. Tel.: + 34 924 289 300; fax: + 34 924 271149.


E-mail address: vgomez@unex.es (V. Gmez-Serrano).
0378-3820/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.fuproc.2010.07.018

structure. Research into the textural characteristics, surface chemistry and


oxidation of AC has been recently reviewed and interpreted by Daud and
Houshamnd [46]. It was concluded by these authors that information
reported in literature about the changes in surface chemistry and textural
characteristics of AC by treatment with a specied oxidant is not similar,
and in some cases is in conict. The possible inconsistencies were
attributed to the diversity of used ACs, oxidants and oxidation methods
and procedures, and analysis methods [46]. In fact, most of the numerous
studies carried out so far on this research topic have been conducted using
a different AC/oxidant system. Nitric acid has been probably the most
used oxidant, because oxidizing specications can be controlled by concentration and temperature [47]. Frequently, not only a single oxidant but
more oxidants have been used in the treatment of a given AC [46,4856].
However, the use of substances in a different gas/liquid phase has received
less attention [5762]. The dry and wet oxidation treatments of AC as a
rule were performed using O2(air) and HNO3 [5862] and rarely also with
H2O2 [57].
Here, using cherry stones as precursor, an AC is prepared by the
method of physical activation in CO2 and the resulting product is
oxidized in O2(air) and in O3 and with HNO3 and H2O2 aqueous
solutions. The analysis of both the virgin AC and AC-derived oxidized
products is carried using FT-IR spectroscopy, Boehm method, and pH
of the point of zero charge for the acidicbasic surface active sites and
gas adsorption (N2, 196 C) and mercury porosimetry for the porous

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

structure. As far as to our knowledge, this is the rst comparison


attempt concerned with the effects of the oxidation of the same AC
with O3 and with oxidants such as O2(air), HNO3 and H2O2 on the
surface chemistry and porous structure.
2. Experimental

1769

Table 1
Methods of oxidation of AC. Values of burn-off percentage. Sample codes.
Oxidant

Flow/L h1

Concentration

T/C

t/h

Burn-off %

Codes

O2(air)
O3
HNO3
H2O2

20
20

Air
1.25 wt.%
5N
5N

300
100
9095
20

24
1
10
10

15.7
1.3
1.7
1.1

ACO2
ACO3
ACHNO3
ACH2O2

2.1. Raw material


Cherry stones (CS, hereafter), a waste generated by the food
industry in the production of brandy and jam, were used here. As
provided by the Agrupacin de Cooperativas Valle del Jerte (Cceres
province, Spain), CS were thoroughly washed with distilled water
under pressure, air- and oven-dried, ground, and sized, the fraction of
particle sizes between 1.6 and 2.0 mm being chosen for subsequent
studies.
2.2. Preparation of activated carbon (AC)
A vertical electrical furnace, provided with a system for temperature control, and a steel reactor were used in the preparation of AC by
the method of physical activation. In the carbonization stage, about
60 g of CS was heated at 20 C min1 from ambient temperature to
900 C in owing N2 (ow = 20 L h1). The soaking time at 900 C
was 2 h. In the activation stage, about 15 g of obtained carbonized
product (C900) was heated at 20 C min1 from room temperature to
850 C in N2 (ow = 20 L h1). Once this temperature was reached,
the N2 atmosphere was switched to the CO2 atmosphere (ow =
20 L h1), which was held for the selected residence time of 2 h.
After that, the furnace was cooled down to room temperature in N2
atmosphere and the resulting product (designated as AC) was stored
in a CaCl2 containing desiccator and weighed.
2.3. Oxidation of AC
2.3.1. Dry
Approximately 10 g of AC was loaded on the steel reactor and
heated at 10 C min1 from room temperature to the desired
oxidation temperature in N2 (ow = 20 L h1). Next, the N2 stream
was replaced with the oxidative stream (ow = 20 L h1) and the
treatment was prolonged at 300 C for 24 h in air and at 100 C for 1 h
in O3-air. After that, the system was cooled down to room temperature in N2 (ow = 20 L h1) and the resulting products were
weighed and stored. A wished concentration of ozone (1.25 wt.%) in
the feeder gas was obtained by xing the voltage in the ozone
generator, Constrema SLO. Ozone was analyzed in the gas stream with
a GM-19 Anseros Ozomat analyzer at 254 nm.
2.3.2. Wet
About 1 g of AC was oxidized with 100 mL of 5 N HNO3 and H2O2
aqueous solutions for 10 h. Whereas the H2O2 treatment was carried
out at room temperature, in the case of the HNO3 treatment the solid
and liquid (oxidizing solution) containing system in continuous
agitation was maintained under a reux at boiling for such a long
time. After each oxidation treatment, the supernatant was separated
by ltration and the residual solid was washed thoroughly with
distilled water, oven-dried at 110 C for 12 h, and stored in glass
bottles. The methods used in the oxidation of AC are summarized in
Table 1, which also lists the sample codes.
2.4. Analyses
2.4.1. Surface chemistry
The surface chemistry of AC and its oxidation products was
analyzed by FT-IR spectroscopy [63,64], Boehm method [65], and pH
of the point of zero charge (pHpzc) [66]. The FT-IR spectra were

recorded on a Perkin-Elmer 1720 FT-IR spectrometer in the 4000


400 cm1 wave number range, 40 scans being taken at a 2-cm1
resolution. Pellets were prepared by mixing powdered sample and
KBr (Merck, for spectroscopy) at a sample/KBr weight ratio of 1:500.
The mixtures were compacted at 10 tonnes/cm2 for 10 min, using a
Perkin-Elmer hydraulic press. A pellet with the same mass as the
carbon containing pellets was prepared from KBr and its spectrum
was recorded and used as background spectrum.
As far as the Boehm method is concerned, the titration of various
acidic oxygen surface groups was effected using 0.05 M NaHCO3 for
carboxylic groups and carboxylic acid anhydrides, 0.05 M Na2CO3 for
lactones and lactols, 0.05 M NaOH for phenolic hydroxyl groups, and
0.25 M NaOH or 0.1 M NaOC2H5 for acidic carbonyl groups. Corrections were introduced in measured titration volumes by allowing that
a base of given strength is able to neutralize those surface functional
groups which are more acidic. On the other hand, a 0.05 M HCl
solution was used for measuring the total content of basic groups [67].
pHpzc was measured using 0.1 M NaNO3 aqueous solutions at pH 3, 5
or 11. These pH values were xed by adding a HNO3 or NaOH aqueous
solution.
2.4.2. Porous structure
The textural characterization of the samples was carried out by gas
adsorption (N2, 196 C) and mercury porosimetry. The N2 isotherms
were determined in a Quantachrome equipment, Autosorb-1. Approximately 0.10 g of the sample was out-gassed at 250 C overnight
under a pressure lower than 103 Torr, prior to the N2 adsorption
measurements. The experiments of mercury intrusion were carried
out in a Quantachrome porosimeter, Autoscan-60, using about 0.3 g of
the sample.
From the N2 adsorption isotherms, the specic surface area of the
samples was evaluated by the BET method (SBET) [68]. Also, the
external surface area (Sex) and micropore volume (Vmi-AI) were
calculated by the s method [69]. The mesopore volume (Vme-AI) was
derived from the adsorption isotherm by subtracting the volume of N2
adsorbed at P/P0 = 0.1 from that at P/P0 = 0.95 (i.e. the total volume of
N2 adsorbed in micro- and mesopores). Vmi-AI and Vme-AI were
expressed as liquid volumes. On the other hand, from the values of
cumulative pore volume (Vcu) against pore radius (r) (mercury
porosimetry) the macropore volume, Vma-P = Vcu (at r = 25 nm), and
the mesopore volume, Vme-P = Vcu (at r b 2 nm) Vma, were obtained.
The total pore volume was calculated as VT = Vmi-AI + Vme-P + Vma-P.
3. Results and discussion
3.1. Oxidation treatments
3.1.1. Burn-off percentage
In Table 1 it is seen that the burn-off percentage of the oxidation
process of AC is higher than zero for all the oxidizing agents.
Accordingly, of both opposite effects on the mass of a sample, i.e. the
consumption of carbon atoms by gasication of AC and the formation
of oxygen surface groups on the material, the former was the
dominant one. In connection with the burn-off percentages it should
be rst kept in mind that the mass decrease produced by AC oxidation
was likely inuenced by the mass of the sample and the reaction
temperature and that these process variables depended on the

1770

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

oxidant (see Table 1). Thus, such a mass was much smaller for H2O2
and HNO3 than for O2(air) and O3 and the oxidation temperature was
higher for HNO3 than for H2O2. This is so because in a series of
exploratory experiments, which were carried out at rst by using the
same mass of AC (i.e. 10 g) for the four oxidizing agents and by
effecting the treatments with the H2O2 and HNO3 solutions at 20 C, it
was observed that the burn-off percentage was very low for both wet
oxidation treatments (i.e. 0.52, H2O2 and 0.94, HNO3) and also that the
treatment with HNO3 was not effective to form oxygen surface groups
in AC, as shown by the FT-IR spectra registered for the resulting
products. These results advised to lower the mass of the sample used
in the wet oxidations and to raise the temperature in the case of the
treatment with the HNO3 solution. In addition, by taking into account
the strong thermal effect on the decomposition rate of H2O2 to H2O
and O2 in aqueous solution (i.e. the rate of the aforesaid process is
roughly doubled by every 10 C increase) and the results obtained in
previous studies on the oxidation of carbonaceous materials in O2(air)
[70] and O3 [63], the heating conditions for the oxidation of AC with
the various chemical agents were chosen.
In accordance with the molecule/ion size of the oxidants (i.e. O2 and
O3 are signicantly smaller chemical species than H2O2 and NO
3 [71])
and with the fact that capillarity phenomena likely occurred when the
aqueous solutions were brought into contact with AC, the diffusion of
the oxidizing agent in pores of AC and the associated degree of
gasication of carbon atoms should be greater, and hence the burn-off
percentage should be higher, for a dry oxidation process than for a wet
oxidation process. Moreover, a greater mass loss was expected with the
decrease in the mass of the sample and with the increase in the
oxidation temperature. As seen in Table 1, the burn-off percentage is
much higher for the oxidation of AC in O2(air) than in O3 and with the
HNO3 and H2O2 solutions. Also, it varies by HNO3 N O3 N H2O2. Accordingly, the higher burn-off percentage for HNO3 than for O3 suggests that
factors other than the mass of the sample and the temperature
inuenced the decrease produced in the mass of AC due to its oxidation.
The relatively high burn-off percentage obtained for O2(air) is
consistent with the heating conditions, i.e. 300 C and 24 h. Thus, as
reported elsewhere [70], the mass of a carbonized product prepared
from CS at 600 C progressively decreased with the increase in
maximum treatment temperature when the material was heated at
250325 C for 24 h in O2(air). At 350 C the initial mass of the sample
was almost entirely gasied-off, yielding ashes [70]. With regard to the
degree of AC gasication by O2(air) action it should be also borne in
mind that the O2 molecule possesses a small size and is linear in shape
and that as a result its access to narrow porosity of AC should be
facilitated, as compared to the rest of the oxidizing agents. In practice, O2
should access even to smaller size pores than those created by partial
gasication of the carbonized product in CO2 during the preparation of
AC (i.e. pores originated as a result of the pyrolysis suffered by CS in the
carbonization stage) on account of the smaller molecular size of O2
(internuclear distance OO, 1.207 ) than of CO2 (internuclear distance
OCO, 2.326 ) [71]. However, the diffusion of O2(air) in pores of AC
was in all likelihood hindered by N2 present in the air stream. In fact, the
N2 molecule (internuclear distance NN, 1.098 ) [71] is smaller than
the O2 molecule and the presence of these molecules in the air stream
was much greater for N2 than for O2. Furthermore, prevention of the
oxidizing action of O2 by N2 should not be ruled out.
The signicantly higher burn-off percentage for HNO3 than for O3
and chiey H2O2 (Table 1) may be surprising since, as an oxidant,
HNO3 is a much less powerful chemical agent than H2O2 and especially
O3 (i.e. E0 at acid pH is equal to + 0.803 V for NO
3 /N2O4, +1.776 V for
H2O2/H2O, and + 2.07 V for O3/O2) [72]. Nevertheless, it should be
taken into account that the oxidation power of the HNO3 solution was
stronger with the increase in temperature during the treatment of AC
because of the larger degree of HNO3 dissociation. Furthermore, this
process was favored by the simultaneous adsorption of the H3O+ ion
on basal planes of AC [73] and by the possible participation of the H3O+

ion in the neutralization of basic sites of AC as these processes should


tend to decrease the presence of H3O+ ions in the bulk of the HNO3
solution. Moreover, under heating conditions, the oxidizing action of
HNO3 also increased as a result of the higher uidity and therefore
greater diffusion of the acid solution in pores of AC. As far as O3 is
concerned, its oxidizing action was likely markedly inuenced by its
thermal stability and internal diffusion in AC. Thus, it has been
previously reported that O3 decomposes above 50 C and that the
process is catalyzed by AC. Also, the catalytic properties of AC in the
decomposition of O3 depend on factors such as the surface area and
volume of pores larger than 3.5 nm [74] or volume of macropores [75].
For H2O2, as compared to HNO3 and O3, not only the decomposition in
aqueous solution but also the larger molecular size could detrimentally
affect its oxidizing action. As a guide, it should be mentioned here that
the OO interatomic distance is 1.28 for O3 and 1.48 for H2O2 and
that the NO interatomic distance is 1.41 for HNO3 [71].
3.2. Surface chemistry
The surface chemistry and porous structure of AC are the two main
factors that control the behavior of this carbon material in adsorption
and catalysis processes. Catalysis on AC surface may play an
outstanding role in oxidation processes of AC because of the oxidant
decomposition. Specically O3 decomposition, among other factors, is
favored by the basicity of the AC surface [75]. It also holds for O3
decomposition in water [76]. For H2O2 decomposition, the driving role
of the chemical nature of the carbon surface in governing the catalytic
action was conrmed by Khalil et al. [77]. Furthermore, the decomposition process occurred to an extent independent of the porosity characteristics of the carbon.
3.2.1. Infrared spectroscopy
The FT-IR spectra registered for AC and AC oxidation products
(Fig. 1) display absorption bands centered at 3500, 17201650, and
12001120 cm1, which are ascribable in turn to stretching () (OH,
C O and CO) vibrations. The bands at 3500 and 12001220 cm1 are

Fig. 1. FT-IR spectra of AC and oxidized products.

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

compatible with the presence of hydroxyl groups in the samples.


According to literature sources [78], the aromatic hydroxyl groups
adsorb infrared radiation at 12601180 cm1 and the aliphatic
hydroxyl groups at 12101000 cm1. The absorption of energy in
the spectral region between 1310 and 1000 cm1 may be also caused
by ether type structures as the (COC) vibration for C COC-al
occurs between 1225 and 1180 cm1. Since ether structures are more
stable thermally than other oxygen surface functionalities [79] and AC
was prepared by heating C900 at 850 C, they could be present in AC
and remain unaltered in the resultant products after the oxidation
treatments. The location of a broad band at 3500 cm1 suggests that
the OH groups are found associated by hydrogen bond [78]. On the
other hand, the absorption of radiation between 1720 and 1650 cm1
was likely caused by (C O) vibrations in oxygen surface groups such
as lactones, carboxylic acid, and quinone. The quinone group usually
absorbs energy at 1700 cm1. However, the corresponding band
may shift to lower frequencies because of insaturation and
increase in conjugation with C C bonds [78,80]. In fact, the band due to
(C C) vibrations in the carboncarbon double bond of olenic structures registers commonly as well at close frequencies. The FT-IR
spectra in Fig. 1 show that the concentration of the oxygen surface
groups absorbing infrared radiation around 1700 cm1 is higher for
ACO3 and ACHNO3, whereas their variety is greater for ACO2 and
ACH2O2.
3.2.2. Boehm method and pHpzc
From the standpoint of the chemical nature of its surface (see data
in Table 2), AC is characterized by possessing low contents of lactones
and carboxylic acid groups. However, the content of hydroxyl groups
is relatively greater in the material. TAG is 138 eq/g for AC, which is
markedly lower than its TBG (i.e. 385 eq/g) and the surface of AC is
therefore basic in character. In fact, pHpzc is equal to 8.8 for AC. The
basicity of carbon surfaces has been associated with a chromene-like
structure, -prone-like structures, and aromatic electrons [47,65]. In
the case of a pyrone-type structure, its strength as a base increases
strongly when the carbonyl group and the ring oxygen are distributed
on polycyclic aromatic compounds [81]. The basic character of the AC
surface is worth noting as carbon basic sites are required for catalysis
reactions in the gas phase such as NO, SO2 or H2S oxidation and in the
liquid phase with H2O2, O3 and wet air [82].
The products of AC oxidation not only contain basic surface sites
but also acidic oxygen surface groups. As compared to AC, the
contents of the various acidic groups are substantially higher for AC
O2. It is so in particular for the lactones and carbonyl groups. As a
result TAG also increases and approaches TBG, pHpzc being equal to
6.8. Concerning ACO3, its TAG is much higher than for AC and ACO2.
Notice the great increase produced in the content of carboxylic acid
groups for ACO3. Since TAG becomes higher than TBG, pHpzc is 4.3 for
ACO3. When a commercial AC was ozonated for 10 and 120 min, its
pHpzc decreased from 8.8 to 6 and 2 respectively [83]. The fact that
Table 2
Surface chemistry of AC and oxidized products. Content of acidicbasic groups (eq/g)
and pHpzc values.
Sample

CGa

Lb

HGc

CaGd

TAGe

TBGf

pHpzcg

AC
ACO2
ACO3
ACHNO3
ACH2O2

16
21
487
639
76

6
27
14
72
49

74
84
98
86
93

42
92
63
69
82

138
224
662
866
300

385
344
302
175
327

8.8
6.7
4.3
3.5
5.3

a
b
c
d
e
f
g

Carboxylic groups.
Lactones.
Hydroxyl groups.
Carbonyl groups.
Total acidic groups.
Total basic groups.
pH of the point zero charge.

1771

TAG can be greatly increased by treatment of AC with O3 must be


highlighted as acidic oxygen surface groups take part in catalysis
reactions of oxidative dehydrogenation, dehydration and dehydrogenation of alcohols, and NOx reduction [82].
Wet oxidation of AC with HNO3 in aqueous solution has proved to
be the most effective treatment to form acidic oxygen surface groups
as the contents of all the groups analyzed in the present study
increased. The effect was much stronger for the carboxylic acid
groups, their content being as high as 639 eq/g in ACHNO3. This
content of oxygen groups is markedly higher than even for ACO3,
which is worth mentioning. As a peculiar feature, the oxidation of AC
with HNO3 is the only treatment that produces an important TBG
decrease. It was expected as HNO3 in aqueous solution is a strong
mineral acid which may then neutralize basic sites of AC. In fact, pHpzc
is as low as 3.5 for CAHNO3. pHpzc values of 3.2 [56] and 2.7 [84] have
been measured previously for commercial ACs oxidized with 5 and
6 mol L1 HNO3 solutions at boiling temperature, respectively. The
surface of the starting ACs, as for the AC used here, was also basic in
character, their pHpzc being in turn 10.1 and 9.7.
In contrast to the treatments of AC with O3 and HNO3, H2O2 did not
enhance the creation, in special, of some acidic oxygen surface group.
With H2O2, the contents of all analyzed oxygen surface groups
increased. Accordingly, H2O2 behaved rather similarly to O2(air).
Nevertheless, it should be noted that the changes produced in the
surface chemistry of AC are somewhat stronger for H2O2 than for O2
(air) as the increase in TAG and the decrease in TBG are signicantly
greater for H2O2. It leads to a signicantly lower pHpzc for ACH2O2.
For this product, however, TAG is much lower and pHpzc is noticeably
higher than for ACO3 and CAHNO3. The pHpzc of 5.3 for ACH2O2 is
in line with the value of 5.2 previously reported in literature for an AC
oxidized also with 10 mol L1 H2O2 solution at room temperature.
When the treatment was effected at 90 C, pHpzc was 4.1 for the
resulting H2O2-oxidized product [56].
In brief, the main factor in controlling the changes produced in the
surface chemistry of AC as a result of its oxidation is the oxidizing
agent rather than whether the process was carried with the substance
in dry phase or in wet phase. The O3 and HNO3 treatments of AC are
the most effective to form acidic oxygen surface groups. The total
content of such groups varies by HNO3 N O3 N H2O2 N O2(air), whereas
for pHpzc the order is HNO3 b O3 b H2O2 b O2(air).
An in-depth literature survey embracing previous studies dealing
with dry and wet oxidation treatments of AC and with their effects on
the surface chemistry and porous structure reveals that most of such
studies were conducted using O2(gas) and HNO3 in aqueous solution
as the oxidizing agents [5762]. Furthermore, they were mostly aimed
at identifying the oxygen surface groups created by AC oxidation,
rather than at measuring their concentration in the starting AC and
AC-oxidized products. As an exception to the rule, the oxygen surface
groups were quantied by temperature programmed desorption
(TPD) and X-ray photoelectron spectroscopy (PXS) by Lemus-Yegres
et al. [61]. From the results obtained in the aforesaid studies it follows
that as a result of the treatments of ACs with both O2(g) and HNO3 in
aqueous solution a variety of oxygen surface groups is formed.
Usually, together with other functional groups, the presence of
phenolic groups and carbonylic groups has been detected in the
case of the O2-oxidized ACs, while carboxylic acid groups and lactone
groups have been found in HNO3-oxidized ACs. The concentration of
carboxylic acid groups was much higher in the HNO3-treated AC and,
oppositely, very low in the O2(air)-treated AC [61]. Therefore, these
results are in line with those obtained in the present study by using
the same couple of oxidizing agents in the treatment of AC.
Ozone in the gas phase has been used as an oxidant in the
treatment of ACs [83,85] and of other carbonaceous materials such as
chars [63,8688], amorphous carbon [89], glassy carbon [90], and so
on. Using a commercial AC, the oxidation with O3 led to the formation
of carboxylic groups, lactone groups, and anhydride groups [83].

1772

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

Jaramillo et al. [85], when investigating the thermal stability of oxygen


surface groups in two ACs prepared from cherry stones by CO2 and
steam activations and in their O3-treated products, have found more
recently that by O3 treatment of the ACs carboxylic groups were
formed to a much larger extent than lactone groups, hydroxyl groups
and carbonyl groups.
In wet oxidation of ACs, H2O2 has been widely used together with
HNO3 [50,52,84,9197]. In addition, (NH4)2S2O8 has also been
frequently employed as an oxidizing agent [50,91,95]. The results
obtained in these works show that H2O2 is a much lesser effective
oxidant than HNO3, concerning the number of oxygen surface groups
created and the extent to which some of them are generated. Using
H2O2, HNO3 and (NH4)2S2O8 in the oxidation of a series of ACs with a
different degree of activation, higher total oxygen content and surface
acidity were obtained for HNO3 [50]. On the contrary, it has also been
reported that the surface chemistry of a commercial AC was not
signicantly affected by treatment with H2O2 at room temperature
and by using a low solution concentration [52]. Furthermore, the
concentration of the oxygen surface groups that desorbed by evolving
CO even decreased [93]. The effectiveness of the H2O2 treatment
increased with the increase in the concentration of the H2O2 solution
[52,84]. Thus, as compared to the starting AC, a higher amount of
oxygenated groups, mainly carboxylic ones, was created when a 10 M
H2O2 solution was used in its oxidation [84]. Also, a rise in the
oxidation temperature favorably inuenced the creation of oxygen
surface groups by H2O2 oxidation, as shown by the higher concentration of lactone groups in an AC oxidation product with H2O2 at
100 C than in the HNO3-treated product at the same temperature.
Furthermore, the concentration of phenolic groups was similar in both
samples of oxidized AC. DTG results obtained in an earlier study on the
thermal behavior of H2O2-treated AC [98] showed the occurrence of
an effect of weight loss centered at about 285 C and also of a marked
change in the mass of the sample between 500 and 700 C. The lowand high-temperature weight losses may be associated with the
thermal decomposition of carboxylic acid groups and of phenolic
groups and carbonyl groups. Moreover, in view of the values of
percentage mass variation as a function of the temperature interval, it
should be noted that the formation of the former groups was more
favorable with the rise of the outgassing temperature of AC to 350 C
prior to oxidation and that of the latter groups with the incorporation
of the Fe2+ ion to the H2O2 solution. In connection with the changes
produced in the surface chemistry of AC by H2O2 oxidation, it has also
been suggested that carbonyl groups are formed from hydroxyl
groups [99].

Fig. 2. N2 adsorption isotherms at 77 K for AC and oxidized products.

N2 adsorption isotherm and from data of mercury porosimetry are


compiled in Table 3. It can be seen that AC chiey contains micro- and
macropores, the mesopore content being much lower. As expected,
the surface area is largely internal for AC. Thus, the percentages of
external and internal surface area are 5.5 and 94.5, respectively.
3.3.2. Oxidized products
From Figs. 2 and 3 and the textural data obtained for the samples in
Table 3, it follows that the oxidation of AC originated changes in its
porous structure which depend on the region of porosity and chemical
agent. Microporosity decreased for all oxidizing agents, the effect
being stronger by HNO3 N O2(air) N O3 N H2O2. The percentage of
microporosity loss (PML), as estimated by the expression PML =
{[Vmi-AI(AC) Vmi-AI(OP)] / Vmi-AI(AC)} 100 (OP means oxidized
product), ranges between 43.3% for ACHNO3 and 6.7% for ACO3.
The strong effect of microporosity loss caused by the oxidation of AC
with HNO3 is worth mentioning. The decrease in the volume of pores
available to N2 adsorption at 196 C has been generally related to
the formation of oxygen groups at the entrance and/or on the walls of
micropores and by the possible destruction of the pore walls and its
collapse when oxygenated terminal groups are created [56,60,101
103]. In view of the opening in the knee of the N2 isotherms and of the
signicantly higher Vme-AI for all oxidized samples, except for CAO2,

3.3. Porous structure


3.3.1. AC
The N2 adsorption isotherm measured for AC (Fig. 2) is a typical
type I isotherm of the BDDT classication system [100]. Accordingly,
as far as the information on the porous structure derived from such an
isotherm is concerned, AC is as an essentially microporous solid.
Furthermore, the high adsorption of N2 at very low relative P/P0
values together with the little pronounced isotherm knee is indicative
of the presence of narrow microporosity in the material. Moreover,
the amount adsorbed of N2 increases rst slightly with the increase in
P/P0 to 0.9 and then more markedly between P/P0 = 0.9 and 1.0. This
evolution of the adsorption of N2 is compatible to the presence in AC
of a heterogeneous mesoporosity made up mainly of wide mesopores.
On the other hand, the plot of the cumulative pore volume against
pore diameter (mercury porosimetry, Fig. 3) shows that in the regions
of intermediate and large size pores AC mostly contains wide mesopores and narrow macropores and that the width of these macropores
ranges between 200 and 600 nm. The contents of a great part of
narrower mesopores and of wider macropores in AC are of little
signicance. The values of the textural parameters obtained from the

Fig. 3. Curves of cumulative pore volume against pore diameter (mercury porosimetry)
for AC and oxidized products.

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

1773

Table 3
Textural parameters of AC and oxidized products.
Sample

Sext/m2 g1

SBET/m2 g1

Vmi-AI/cm3 g1

Vme-AI/cm3 g1

Vme-P/cm3 g1

Vma-P/cm3 g1

VT/cm3 g1

AC
ACO2
ACO3
ACHNO3
ACH2O2

35
37
46
50
50

604
469
603
399
591

0.30
0.21
0.28
0.17
0.26

0.069
0.056
0.105
0.090
0.087

0.055
0.062
0.073
0.097
0.084

0.34
0.34
0.36
0.35
0.35

0.70
0.61
0.71
0.62
0.69

it seems also that micropores initially present in AC as a rule widened


and transformed into large size micro- and mesopores. Concerning
CAO2, notice that the isotherm becomes more parallel to the abscissa
axis and hence Vme-AI is noticeably lower for this sample than for AC.
The curves of cumulative pore volume against pore diameter (Fig. 3)
clearly show that the content of wide mesopores is higher for AC than
for ACO2, whereas the opposite applies to the content of narrow
mesopores. This may account for the signicantly higher Vme-P for AC
O2 than for AC. Perhaps, mercury could enter a fraction of pores
present in ACO2 which underwent an enlargement because of the
application of very high pressures in the mercury porosimeter. Such
pores might possess thin walls as a consequence of the high degree of
gasication of AC in O2(air) during the preparation of ACO2 and
hence they were crushed during the subsequent porosimetry run. On
the other hand, this interpretation of the results would be also compatible with lower Vme-AI for ACO2 than for AC. Mesoporosity is
better developed in ACO3, ACHNO3 and ACH2O2 than in AC. For
ACO3, the content of narrow mesopores also decreases (Fig. 3).
Practically, macroporosity as a rule was not affected by the oxidation of AC. The most important change in Vma-P appears for ACO3, as
it is 0.36 cm3/g for this sample and 0.34 cm3/g for AC. Furthermore,
the derivative curves of mercury intrusion volume against pore
diameter obtained for AC and oxidized products (i.e. they are omitted
for the sake of brevity) show the presence of a strong maximum
centered at a value of the pore diameter of about 525 nm for ACH2O2
and ACO3 and of about 375 nm for AC, ACO2 and ACHNO3. In
addition, the curve obtained for ACO3 exhibits another maximum
located at about 265 nm. Accordingly, the presence of wider macropores in ACH2O2 and ACO3 and of a more heterogeneous pore size
distribution in the range of macropores for ACO3 becomes apparent.
On the other hand, in Table 3 it is seen that VT noticeably decreases for
ACO2 and ACHNO3. For ACO3 and ACH2O2, however, the decrease
of VT is of little signicance.
In brief, concerning the effect of the oxidation of AC on the porous
structure, the behavior shown by O3 and H2O2 is rather similar as with
both oxidizing agents a widening of narrower pores occurs, which
causes microporosity loss and mesoporosity development. With HNO3
the decrease produced in microporosity is much greater. Although to a
signicantly lesser extent, it also applies to O2(air). In this case,
however, there is no pore enlargement but simply micropore loss. By
keeping in mind the great mass loss produced in the treatment of AC
with O2(air), the occurrence of a rapid formation of transient oxygen
surface groups leading eventually to gasication of pore walls,
without yielding larger size pores, is possible.
Results obtained in previous studies on the oxidation of AC reveal
that the changes produced in the porous structure of the material by
dry and wet treatments are strongly dependent on the oxidizing agent
and heating conditions and, for processes occurring in wet phase, also
on the concentration of the oxidant solution. Using O2(air), microporosity was not substantially modied by treatment at 300 C for
90 min [57]. Also, the changes produced in textural properties of the
carbon were only slight at 350 C for 3 h [61]. However, BET surface
area signicantly developed at 420 C [60]. For HNO3, PML was as high
as 69 for the oxidation of a commercial AC with 5 M acid solution at
the boiling temperature for 6 h. In spite of this strong microporosity
reduction, it should also be noted that PML was as low as 9.8 [57] and

3.8 [84] for ACs reacted with HNO3 under a solution reux and that, as
a last resort, microporosity modications were only small for HNO3
treatments at room temperature [57,61]. With HNO3, also, the
reduction of microporosity was greatly favored with the increase in
oxidation time and concentration of the HNO3. PML was 18.4 and 56.7
for 4 and 48 h treatment times [104] and 16.3 and 69.2 when 2 and
5 M HNO3 solutions were used. It was also observed that the
detrimental effect of HNO3 concentration on microporosity was
dependent on treatment time [62]. In the case of O3 treatments,
PMLs ranging between 0.42 and 37.3 were obtained for 10120 min
ozonation [83]. For the treatment of two CO2 and steam-activated
carbons with O3, PML was 7.0 and 0.8, respectively [85]. Finally, the
changes caused in the surface area and micropore volume of ACs by
H2O2 oxidation were generally of little signicance (i.e. PML b ~10)
[52,57,84,9193], becoming more important when the treatment was
effected at temperatures higher than room temperature [55,56] or
during prolonged time periods [53]. Microporosity decreased with the
rise of the oxidation temperature [55,56], whereas the surface area
developed with a long reaction time [53]. With the increase of both
process variables, the opposite effects on the porous structure seemed
to balance [96]. An increase in the concentration of the H2O2 solution
had no signicant effect on microporosity [95]. In brief, by taking
mainly into consideration the heating conditions used in the oxidation
of different ACs, it may be concluded that the results obtained in the
present study on textural changes originated in AC by oxidation with
HNO3, O3 and H2O2 are well in agreement with those previously
reported in literature. The strong detrimental effect of the oxidation of
AC in O2(air) at 300 C for 24 h on microporosity and surface area is
attributable to the very long reaction time as the treatment was
performed at a relatively low temperature.
4. Conclusions
From the above results the following conclusions may be drawn:
(1) The oxidation of AC (i.e. an activated carbon prepared from cherry
stones by CO2 activation) with dry O2(air) and O3 and with wet
HNO3 and H2O2 increases the content of acidic oxygen surface
groups, the most effective chemical agents being HNO3 and O3.
(2) The treatments of AC with O3 and HNO3 result in the formation
of mostly carboxylic acid groups.
(3) The content of basic sites greatly decreases for HNO3.
(4) The oxidation treatments of AC cause a decrease of microporosity which is more important for O2(air) and especially for
HNO3. However, as a rule, mesoporosity signicantly develops
and macroporosity remains unchanged.
(5) The dry oxidation of AC at 25 C in O3 is the most promising
treatment to form acidic oxygen groups in the carbon surface
without greatly altering its content of basic groups and volume
of micropores.
Acknowledgements
This work has been supported by the Junta de Extremadura
(Spain) through project PDT08-A012. Dr. J. Jaramillo also thanks the
Consejera de Educacin of Junta de Extremadura for providing her a

1774

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775

sabbatical research stay at the Departamento de Ingeniera Qumica of


the University of Extremadura.
References
[1] Z. Huang, X. Liu, K. Li, D. Li, Y. Luo, H. Li, W. Song, L. Chen, Q. Meng, Application of
carbon materials as counter electrodes of dye-sensitized solar cells, Electrochem.
Commun. 9 (2007) 596598.
[2] K. Li, Y. Luo, Z. Yu, M. Deng, D. Li, Q. Meng, Low temperature fabrication of
efcient porous carbon counter electrode for dye-sensitized solar cells,
Electrochem. Commun. 11 (2009) 13461349.
[3] H. Peng, F. Cheng, J. Shi, J. Liang, Z. Tao, J. Chen, High-surface area microporous
carbon as the efcient photocathode of dye-sensitized solar cells, Solid State Sci.
11 (2009) 20512055.
[4] A. Guha, W. Lu, T.A. Zawodzinski Jr., D.A. Schiraldi, Surface-modied carbons as
platinum catalyst support for PEM fuel cells, Carbon 45 (2007) 15061517.
[5] X. Ma, D. Yang, W. Zhou, C. Zhang, X. Pan, L. Xu, M. Wu, J. Ma, Evaluation of
activated carbon adsorbent for fuel cell cathode air ltration, J. Power Sources
175 (2008) 383389.
[6] D. Jiang, B. Li, Granular activated carbon single-chamber microbial fuel cells
(GAC-SCMFCs): a design suitable for large-scale wastewater treatment processes, Biochem. Eng. J. 47 (2009) 3137.
[7] X. Li, Z. Zhu, J. Chen, R. de Marco, A. Dicks, J. Bradley, G. Lu, Surface modication of
carbon fuels for direct carbon fuel cells, J. Power Sources 186 (2009) 19.
[8] G. Zini, R. Marazzi, S. Pedrazzi, P. Tartarini, A solar hydrogen hybrid system with
activated carbon storage, Int. J. Hydrogen Energ. 35 (2010) 49094917.
[9] A.G. Pandolfo, A.F. Hollenkamp, Carbon properties and their role in supercapacitors, J. Power Sources 157 (2006) 1127.
[10] E. Raymundo-Piero, K. Kierzek, J. Machnikowski, F. Bguin, Relationship
between the nanoporous texture of activated carbons and their capacitance
properties in different electrolytes, Carbon 44 (2006) 24982507.
[11] A. Burke, R&D considerations for the performance and application of electrochemical capacitors, Electronchim. Acta 53 (2007) 1081091.
[12] M. Olivares-Marn, J.A. Fernndez, M.J. Lzaro, C. Fernndez-Gonzlez, A. MacasGarca, V. Gmez-Serrano, F. Stoeckli, T.A. Centeno, Cherry stones as precursor of
activated carbons for supercapacitors, Mater. Chem. Phys. 114 (2009) 323327.
[13] A. Caballero, L. Hernn, J. Morales, M. Olivares-Marn, V. Gmez-Serrano,
Suppressing irreversible capacity in low cost disordered carbons for Li-ion
batteries, Electrochem. Solid St. 12 (2009) A167A170.
[14] J.C. Arrebola, A. Caballero, L. Hernn, J. Morales, M. Olivares-Marn, V. GmezSerrano, Improving the performance of biomass-derived carbons in Li-ion
batteries by controlling the lithium insertion process, J. Electrochem. Soc. 157
(2010) A791A797.
[15] T.D. Burchell, Carbon Materials for Advanced Technologies, Pergamon, Amsterdam, 1999.
[16] R.K. Ahluwalia, J.K. Peng JK, Automotive hydrogen storage system using cryoadsorption on activated carbon, Int. J. Hydrogen Energ. 34 (2009) 54765487.
[17] F. Rodrguez-Reinoso, Y. Nakagawa, J. Silvestre-Albero, J.M. Jurez-Galn, M.
Molina-Sabio, Correlation of methane uptake with microporosity and surface
area of chemically activated carbons, Micropor. Mesopor. Mat. 115 (2008)
603608.
[18] M. Namvar-Asl, M. Soltanieh, A. Rashidi, Modeling and preparation of activated
carbon for methane storage II. Neural network modelling and experimental
studies of activated carbon preparation, Energ. Convers. Manage 49 (2008)
24782482.
[19] H. Najibi, A. Chapoy, B. Tohidi, Methane/natural gas storage and delivered
capacity for activated carbons in dry and wet conditions, Fuel 87 (2008) 713.
[20] J. Alcaiz-Monge, D. Lozano-Castell, D. Cazorla-Amors, A. Linares-Solano,
Fundamentals of methane adsorption in microporous carbons, Micropor.
Mesopor. Mat. 124 (2009) 110116.
[21] E. Jurewicz, E. Frackowiak, F. Beguin, Electrochemical storage of hydrogen in
activated carbons, Fuel Process. Technol. 7778 (2002) 415421.
[22] M. Jord-Beneyto, F. Surez-Garca, D. Lozano-Castell, D. Cazorla-Amors, A.
Linares-Solano, Hydrogen storage on chemically activated carbons and carbon
nanomaterials at high pressures, Carbon 45 (2007) 293303.
[23] M.H. Kim, E.K. Lee, J.H. Jun, S.J. Kong, G.H. Han, B.K. Lee, T.-J. Lee, K.J. Yoon,
Hydrogen production by catalytic decomposition of methane over activated
carbons: kinetic study, Int. J. Hydrogen Energ. 29 (2004) 187193.
[24] R. Moliner, I. Suelves, M.J. Lzaro, O. Moreno, Thermocatalytic decomposition of
methane over activated carbons: inuence of textural properties and surface
chemistry, Int. J. Hydrogen Energ. 30 (2005) 293300.
[25] N. Muradov, F. Smith, A. T-Raissi, Catalytic activity of carbons for methane
decomposition reaction, Catal. Today 102103 (2005) 225233.
[26] W. Dai, Y. Zhou, S. Wang, W. Su, Y. Sun, L. Zhou, Desulfurization of transportation
fuels targeting at removal of thiophene/benzothiophene, Fuel Process. Technol.
89 (2008) 749755.
[27] H.S. Bamueh, Single and binary sulfur removal components from model diesel
fuel using granular activated carbon from dates' stones activated by ZnCl2, Appl.
Catal. A: Gen. 365 (2009) 153158.
[28] M. Seredych, J. Lison, U. Jans, T.J. Bandosz, Textural and chemical factors affecting
adsorption capacity of activated carbon in highly efcient desulfurization of
diesel fuel, Carbon 47 (2009) 24912500.
[29] M. Muzic, K. Sertic-Bionda, Z. Gomzi, S. Podolski, S. Telen, Study of diesel fuel
desulfurization by adsorption, Chem. Eng. Res. Des. 88 (2010) 487495.

[30] G. Yu, S. Lu, H. Chen, Z. Zhu, Diesel fuel desulfurization with hydrogen peroxide
promoted by formic acid and catalyzed by activated carbon, Carbon 43 (2005)
22852294.
[31] V. Selvavathi, V. Chidambaram, A. Meenakshisundaram, B. Sairan, B. Sivasankar,
Adsorptive desulfurization of diesel on activated carbon and nickel supported
systems, Catal. Today 141 (2009) 99102.
[32] T.C. Drage, O. Kozynchenko, C. Pevida, M.G. Plaza, F. Rubiera, J.J. Pis, C.E. Snape, S.
Tennison, Developing activated carbon adsorbents for pre-combustion CO2
capture, Energy Procedia 1 (2009) 599605.
[33] D. Cao, J. Wu, Modelling the selectivity of activated carbons for efcient
separation of hydrogen and carbon dioxide, Carbon 43 (2005) 13641370.
[34] T.J. Bandosz, Removal of inorganic gases and VOCs on activated carbon,
Adsorption by Carbons, Elsevier, Amsterdam, 2008.
[35] A. Somy, M.R. Mehmia, H.D. Amrei, A. Ghanizadeh, M. Safari, Adsorption of
carbon dioxide using impregnated activated carbon promoted by zinc, Int. J.
Greenhouse Gas Control 3 (2009) 249253.
[36] M. Pellerano, P. Pr, M. Kacem, A. Delebarre, CO2 capture by adsorption on
activated carbons using pressure modulation, Energy Procedia 1 (2009)
647653.
[37] H. Yang, Z. Xu, M. Fan, R. Gupta, R.B. Slimane, A.E. Bland, I. Wright, Progress in
carbon dioxide separation and capture: a review, J. Environ. Sciences 20 (2008)
1427.
[38] D. Lozano-Castell, J. Alcaiz-Monge, M.A. de la Casa-Lillo, D. Cazorla-Amors, A.
Linares-Solano, Advances in the study of methane storage in porous carbonaceous materials, Fuel 81 (2002) 17771803.
[39] A. Silvestre-Albero, J. Silvestre-Albero, A. Seplveda-Escribano, F. RodrguezReinoso, Ethanol removal using activated carbon: effect of porous structure and
surface chemistry, Micropor. Mesopor. Mat. 120 (2009) 6268.
[40] T.J. Bendosz, Effect of pore structure and surface chemistry of virgin activated
carbons on removal of hydrogen sulde, Carbon 37 (1999) 483491.
[41] P. Davini, Adsorption of sulphur dioxide on thermally treated active carbon, Fuel
68 (1989) 145148.
[42] P. Davini, Adsorption and desorption of SO2 on active carbon. The effect of
surface basic groups, Carbon 28 (1990) 565571.
[43] N. Karatepe, I. Orbak, R. Yavuz, A. zyuguran, Sulfur dioxide adsorption by
activated carbons having different textural and chemical properties, Fuel 87
(2008) 32073215.
[44] A. Lisvskii, R. Semiat, C. Aharoni, Adsorption of sulfur dioxide by active carbon
treated by nitric acid: I. Effect of the treatment on adsorption of SO2 and
extractability of the acid formed, Carbon 35 (1997) 16391643.
[45] J.T. Cookson Jr, Adsorption mechanisms: the chemistry of organic adsorption on
activated carbon, in: P.N. Cheremisinoff, F. Ellerbusch (Eds.), Carbon Adsorption
Handbook, Ann Arbor Sci, Ann Arbor, 1980, pp. 241279.
[46] W.M.A.W. Daud, A.H. Houshamnd, Textural characteristics, surface chemistry
and oxidation of activated carbon, J. Natur. Gas Chem. 19 (2010) 267279.
[47] H.P. Boehm, Surface oxides on carbon and their analysis: a critical assessment,
Carbon 40 (2002) 145149.
[48] M. Acedo-Ramos, V. Gomez-Serrano, C. Valenzuela-Calahorro, A.J. LopezPeinado, Oxidation of activated carbon in liquid phase. Study by FT-IR, Spectrosc.
Lett. 26 (1993) 11171137.
[49] C. Moreno-Castilla, M.A. Ferro-Garca, J.P. Joly, I. Bautista-Toledo, F. CarrascoMarn, J. Rivera-Utrilla, Activated carbon surface modications by nitric acid,
hydrogen peroxide, and ammonium peroxydisulfate treatments, Langmuir 11
(1995) 43864392.
[50] C. Moreno-Castilla, M.V. Lpez-Ramn, F. Carrasco-Marn, Changes in surface
chemistry of activated carbons by wet oxidation, Carbon 38 (2000) 19952001.
[51] I.I. Salame, T.J. Bandosz, Surface chemistry of activated carbons: combining the
results of temperature-programmed desorption, Boehm, and potentiometric
titrations, J. Colloid Interface Sci. 240 (2001) 252258.
[52] M.F.R. Pereira, S.F. Soares, J.J.M. rfao, J.L. Figueiredo, Adsorption of dyes on
activated carbons: inuence of surface chemical groups, Carbon 41 (2003)
811821.
[53] Y. Xue, Y. Guo, Z. Zhang, Y. Guo, Y. Wang, G. Lu, The role of surface properties of
activated carbon in the catalytic reduction of NO by carbon, Appl. Surf. Sci. 255
(2008) 25912595.
[54] A.H. El-Sheikh, Effect of oxidation of activated carbon on its enrichment
efciency of metal ions: comparison with oxidized and non-oxidized multiwalled carbon nanotubes, Talanta 75 (2008) 127134.
[55] C.-C. Huang, H.-M. Chen, C.-H. Chen, J.-C. Huang, Effect of surface oxides on
hydrogen storage of activated carbon, Sep. Purif. Technol. 70 (2010) 291295.
[56] X. Song, H. Liu, L. Cheng, Y. Qu, Surface modication of coconut-based activated
carbon by liquid-phase oxidation and its effects on lead ion adsorption,
Desalination 255 (2010) 7883.
[57] M. Molina-Sabio, M.A. Muecas-Vidal, F. Rodriguez-Reinoso, Modication in
porous texture and oxygen surface groups of activated carbons by oxidation,
Stud. Surf. Sci. Catal. 62 (1991) 329339.
[58] J.A.C. Alves, C. Freire, B. de Castro, J.L. Figueiredo, Anchoring of organic molecules
onto activated carbon, Colloids Surf. A 189 (2001) 7584.
[59] V. Strelko Jr., D.J. Malik, M. Streat, Characterisation of the surface of oxidised
carbon adsorbents, Carbon 40 (2002) 95104.
[60] Y.H. Li, C.W. Lee, B.K. Gullet, Importance of activated carbon's oxygen surface
functional groups on elemental mercury adsorption, Fuel 82 (2003) 451457.
[61] L.J. Lemus-Yegres, I. Such-Basez, M.C. Romn-Martnez, C. Salinas-Martnez de
Lecea, Catalytic properties of a RH-diamine complex anchored on activated
carbon: effect of different surface oxygen groups, Appl. Catal. A: Gen. 331 (2007)
2633.

J. Jaramillo et al. / Fuel Processing Technology 91 (2010) 17681775


[62] G.G. Stavropoulos, P. Samaras, G.P. Sakellaropoulos, Effects of activated carbon
modication on porosity, surface structure and phenol adsorption, J. Hazard.
Mater. 151 (2008) 414421.
[63] V. Gmez-Serrano, P.M. lvarez, J. Jaramillo, Formation of oxygen complexes by
ozonation of carbonaceous materials prepared from cherry stones: I. Thermal
effects, Carbon 40 (2002) 513522.
[64] V. Gmez-Serrano, M.C. Fernndez-Gonzlez, M. Alexandre-Franco, A. MacasGarca, Optimizing the application of infrared spectroscopy for the study of
carbonaceous materials, Ann. Chim. Sci. Mat. 30 (2005) 7794.
[65] H.P. Boehm, Some aspects of the surface chemistry of carbon blacks and other
carbons, Carbon 32 (1994) 759769.
[66] J.S. Noh, J.A. Schwarz, Estimation of the point of zero charge of simple oxides by
mass titration, J. Colloid Interf. Sci. 130 (1989) 157164.
[67] F. Carrasco-Marn, A. Mueden, T.A. Centeno, F. Stoeckli, C. Moreno-Castilla, Water
adsorption on activated carbons with different degrees of oxidation, J. Chem. Soc.
Faraday Trans. 93 (1997) 22112215.
[68] S. Brunauer, P.H. Emmett, E. Teller, Adsorption of gases in multimolecular layers,
J. Am. Chem. Soc. 60 (1938) 309319.
[69] K.S.W. Sing, Surface Area Determination, Butterworths, London, 1970.
[70] V. Gmez-Serrano, F. Piriz-Almeida, C.J. Durn-Valle, J. Pastor-Villegas, Formation of oxygen by air activation. A study by FT-IR spectroscopy, Carbon 37 (1999)
15171528.
[71] N.N. Greenwood, A. Earnshaw, Chemistry of the Elements, Pergamon, Oxford,
1989.
[72] D.R. Lide (Ed.), CRC Handbook of Chemistry and Physics, 86th ed, Taylor &
Francis, Boca Raton, 2005.
[73] C.A. Leon y Leon, J.M. Solar, V. Calemma, L.R. Radovic, Evidence for the
protonation of basal plane sites on carbon, Carbon 30 (1992) 797811.
[74] P.M. lvarez, F.J. Masa, J. Jaramillo, F.J. Beltrn, V. Gmez-Serrano, Kinetics of
ozone decomposition by granular activated carbon, Ind. Eng. Chem. Res. 47
(2008) 25452553.
[75] J. Rivera-Utrilla, M. Snchez-Polo, Ozonation of 1,3,6-naphthalenetrisulphonic
acid catalysed by activated carbon in aqueous phase, Appl. Catal. B: Environ. 39
(2002) 319329.
[76] P.C.C. Faria, J.J.M. rfao, M.F.R. Pereira, Ozone decomposition in water catalyzed
by activated carbon: inuence of chemical and textural properties, Ind. Eng.
Chem. Res. 45 (2006) 27152721.
[77] L.B. Khalil, B.S. Girgis, T.A.M. Tawk, Decomposition of H2O2 on activated carbon
obtained from olive stones, J. Chem. Technol. Biot. 76 (2001) 11321140.
[78] E. Pretsch, T. Clerc, J. Seibl, W. Simon, Tablas para la Elucidacin Estructural de
Compuestos Orgnicos por Mtodos Espectroscpicos, Alambra, Madrid, 1980.
[79] H. Jankowska, A. Swiatkowski, J. Choma, Active Carbon, Ellis Horwood,
New York, 1991.
[80] V. Gomez-Serrano, J. Pastor-Villegas, A. Perez-Florindo, C.J. Duran-Valle, C.
Valenzuela-Calahorro, FT-IR study of rockrose and of char and activated carbon,
J. Anal. Appl. Pyrol. 36 (1996) 7180.
[81] D. Surez, J.A. Menndez, E. Fuente, M.A. Montes-Morn, Contribution of pyronetype structures to carbon basicity: an ab initio study, Langmuir 15 (1999)
38973904.
[82] J.L. Figueiredo, M.F.R. Pereira, The role of surface chemistry in catalysis with
carbons, Catal. Today 150 (2010) 27.
[83] H. Valds, M. Snchez-Polo, J. Rivera-Utrilla, C.A. Zaror, Effect of ozone treatment
on surface properties of activated carbon, Langmuir 18 (2002) 21112116.
[84] P.C.C. Faria, J.J.M. rfao, M.F.R. Pereira, Adsorption of anionic and cationic dyes on
activated carbons with different surface chemistries, Water Res. 38 (2004)
20432052.

1775

[85] J. Jaramillo, P.M. lvarez, V. Gmez-Serrano, Preparation and ozone-surface


modication of activated carbon. Thermal stability of oxygen surface groups,
Appl. Surf. Sci. 256 (2010) 52325236.
[86] Y. Otake, R.G. Jenkins, Characterization of oxygen-containing surface complexes
created on a microporous carbon by air and nitric acid treatment, Carbon 31
(1993) 109121.
[87] V. Gmez-Serrano, P.M. lvarez, J. Jaramillo, F.J. Beltrn, Formation of oxygen
complexes by ozonation of carbonaceous materials prepared from cherry stones.
II. Kinetic study, Carbon 40 (2002) 523529.
[88] L.A. Langley, D.H. Fairbrother, Effect of wet chemical treatments on the
distribution of surface oxides on carbonaceous materials, Carbon 45 (2007)
4754.
[89] D.B. Mawhinney, J.T. Yates Jr., FTIR study of the oxidation of amorphous carbon
by ozone at 300 Kdirect COOH formation, Carbon 39 (2001) 11671173.
[90] F.J. Lpez-Garzn, M. Domngo-Garca, M. Prez-Mendoza, P.M. Alvarez, V.
Gmez-Serrano, Textural and chemical surface modications produced by some
oxidation treatments of a glassy carbon, Langmuir 19 (2003) 28382844.
[91] B.K. Pradhan, N.K. Sandle, Effect of different oxidizing agent treatments on the
surface properties of activated carbons, Carbon 37 (1999) 13231332.
[92] W. Qiao, Y. Korai, I. Mochida, Y. Hori, T. Maeda, Preparation of an activated
carbon artifact: oxidative modication of coconut shell-based carbon to improve
the strength, Carbon 40 (2002) 351358.
[93] T. Garca, R. Murillo, D. Cazorla-Amors, A.M. Mastral, A. Linares-Solano, Role of
the activated carbon surface chemistry in the adsorption of phenanthrene,
Carbon 42 (2004) 16831689.
[94] N. Zhao, N. Wei, J. Li, Z. Qiao, J. Cui, F. He, Surface properties of chemically
modied activated carbons for adsorption of Cr(VI), Chem. Eng. J. 115 (2005)
133138.
[95] M. Santiago, F. Stber, A. Fortuny, A. Fabregat, J. Font, Modied activated carbon
for catalytic wet air oxidation of phenol, Carbon 43 (2005) 21342145.
[96] M. Takaoka, H. Yokokawa, N. Takeda, The effect of treatment of activated carbon
by H2O2 or HNO3 on the decomposition of pentachlorobenzene, Appl. Catal. B:
Environ. 74 (2007) 179186.
[97] V.Z. Radkevich, T.L. Senko, K. Wilson, L.M. Grishenko, A.N. Zaderko, V.Y. Diyuk,
The inuence of surface functionalization of activated carbon on palladium
dispersion and catalytic activity in hydrogen oxidation, Appl. Catal. A: Gen. 335
(2008) 241251.
[98] V. Gmez-Serrano, M. Acedo-Ramos, A.J. Lpez-Peinado, C. Valenzuela-Calahorro,
Thermogravimetric study of activated carbon oxidized with H2O2, Thermochim.
Acta 254 (1995) 249260.
[99] V. Gmez-Serrano, M. Acedo-Ramos, A.J. Lpez-Peinado, C. ValenzuelaCalahorro, Oxidation of activated carbon by hydrogen peroxide. Study of
surface functional groups by FT-i.r. Fuel 73 (1994) 387395.
[100] S. Brunauer, L.S. Deming, W.S. Deming, E. Teller, On the theory of the van der
Waals adsorption of gases, J. Am. Chem. Soc. 62 (1940) 17231732.
[101] C. Moreno-Castilla, F. Carrasco-Marn, A. Mueden, The creation of acid carbon
surfaces by treatment with (NH4)2S2O8, Carbon 35 (1997) 16191626.
[102] C. Moreno-Castilla, F. Carrasco-Marn, F.J. Maldonado-Hodar, J. Rivera-Utrilla,
Effects of non-oxidant and oxidant acid treatments on the surface properties o
fan activated carbon with very low ash content, Carbon 36 (1998) 145151.
[103] B. Xiao, K.M. Thomas, Adsorption of aqueous metal ions on oxygen and nitrogen
functionalized nanoporous activated carbons, Langmuir 21 (2005) 38923902.
[104] Y.F. Jia, K.M. Thomas, Adsorption of cadmium ions on oxygen surface sites in
activated carbon, Langmuir 16 (2000) 11141122.

Você também pode gostar