Вы находитесь на странице: 1из 7

PERGAMON

Carbon 38 (2000) 19952001

Changes in surface chemistry of activated carbons by wet


oxidation

b , F. Carrasco-Marn
a
C. Moreno-Castilla a , *, M.V. Lopez-Ramon
a

, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain


Departamento de Qumica
Inorganica

, 23071 Jaen
, Spain
Departamento de Qumica
Inorganica
y Organica
, Facultad de Ciencias Experimentales, Universidad de Jaen
Received 18 October 1999; accepted 16 February 2000

Abstract
A series of activated carbons with different degrees of activation were oxidized with H 2 O 2 , (NH 4 ) 2 S 2 O 8 and HNO 3 in
order to introduce different oxygen surface complexes. Changes in the surface chemistry of the activated carbons after their
oxidizing treatments were studied by different techniques including temperature-programmed desorption (TPD), X-ray
photoelectron spectroscopy (XPS), Fourier transformed infrared spectroscopy (FTIR), titrations with HCl and NaOH,
measurements of the pH of the point of zero charge and catalytic dehydration of methanol. Results showed that treatment
with (NH 4 ) 2 S 2 O 8 fixed the lowest amount of both total oxygen and surface acid groups. However, this treatment yielded the
acid groups with the highest acid strength. This could be because it favors fixation of carboxyl groups close to other groups,
such as carbonyl and hydroxyl, which enhances their acidity. 2000 Elsevier Science Ltd. All rights reserved.
Keywords: A. Activated carbon; B. Oxidation; D. Surface oxygen complexes

1. Introduction
The surface chemistry of carbon materials is basically
determined by the acidic and basic character of their
surface, and can be changed by treating them with
oxidizing agents either in the gas phase or in solution.
These treatments fix a certain amount of oxygen surface
complexes such as: carboxyls, lactones, phenols, ketones,
quinones, alcohols and ethers that make the carbon materials more hydrophilic and acidic, decreasing the pH of their
point of zero charge and increasing their surface charge
density [111]. At the same time, the surface basicity has
been found to decrease suggesting that the surface basic
sites are essentially of the Lewis type, associated with p
electron-rich regions found on the basal planes. Thus, an
increase in the oxygen content of the carbon diminishes the
electronic density of the basal planes and, consequently,
reduces the basicity of the carbon surface [1119].
Different oxidizing agents can be used in aqueous
solutions to introduce oxygen surface complexes [12], and
we have shown recently [7,8] that not all of these behave
*Corresponding author. Tel.: 134-958-243-323; fax: 134958-248-526.
E-mail address: cmoreno@ugr.es (C. Moreno-Castilla).

in the same way. Thus, when activated carbons prepared


from almond shells were oxidized with HNO 3 or
(NH 4 ) 2 S 2 O 8 the latter treatment yielded oxidized activated
carbons with stronger acid groups than the former, in spite
of the fewer oxygen surface complexes [7,8]. The aim of
the present work was to gain more insight into the changes
occurring in the surface chemistry of activated carbons
with different degrees of activation, after their wet oxidation with nitric acid, hydrogen peroxide and ammonium
peroxydisulphate. These changes were followed by temperature-programmed desorption (TPD), X-ray photoelectron spectroscopy (XPS), Fourier transformed infrared
spectroscopy (FTIR), titrations with NaOH and HCl,
measurements of the pH of the point zero charge, pH PZC ,
and by the behavior of the oxidized carbons in the
dehydration reaction of methanol.

2. Experimental
Activated carbons were prepared from olive stones. The
raw material supplied by an olive mill was sieved to obtain
a particle size between 1 and 1.5 mm, which was treated
with a diluted H 2 SO 4 solution (10%) and washed with
distilled water till absence of sulphate ions in the washing

0008-6223 / 00 / $ see front matter 2000 Elsevier Science Ltd. All rights reserved.
PII: S0008-6223( 00 )00048-8

1996

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001

water. This material was carbonized in a N 2 flow at 1273


K for 1 h, and after that, steam activated at 1123 K for
different periods of time to obtain various degrees of
activation, following the method explained elsewhere [8].
The activated carbons so prepared will be referred to in the
text as BV followed by a number that indicates the degree
of activation, or percentage of burn-off, during the steam
activation step.
The samples obtained were oxidized with HNO 3 , H 2 O 2
and (NH 4 ) 2 S 2 O 8 . The oxidized samples, or the treatment
followed to obtain them, will be referred to in the text as
N, H or S, respectively. The procedures followed to
oxidize the activated carbons have been explained in detail
elsewhere [7,8]. After the N and S treatments the samples
were washed with distilled water till absence of nitrate and
sulphate ions, respectively, in the washing water.
All samples were characterized by N 2 adsorption at 77
K by applying the BET equation to the N 2 adsorption
isotherm. Mercury porosimetry data were obtained up to a
22
final pressure of 4200 kg cm
using an Austoscan 60
equipment. From this technique, the pore volume of pores
with a diameter between 3.6 and 50 nm, V2 , and the pore
volume of pores with a diameter greater than 50 nm were
obtained.
Surface chemistry of activated carbons was studied by
TPD, XPS, FTIR, titration with NaOH and HCl and pH PZC
measurements. TPD experiments were carried out by
heating the samples up to 1273 K in He flow at a heating
rate of 20 K min 21 and recording the amounts of CO and
CO 2 evolved with a quadrupole mass spectrometer, from
Balzers, model Thermocube, as a function of temperature
as described elsewhere [20].
XPS measurements were made with a Physical Electronic 5700 equipment with a Mg K a X-ray excitation
source (hn 51253.6 eV) and hemispherical electron analyzer. Prior to the analysis, the samples were heated in situ
at 393 K for 1 h under high vacuum and introduced in the
analysis chamber without any contact with the atmosphere.
The residual pressure in the analysis chamber was maintained below 10 29 Torr during data acquisition. Survey
and multiregion spectra were recorded at O 1s and C 1s
photoelectron peaks. The atomic concentrations were
calculated from the photoelectron peak areas, using Shirley
background subtraction [21] and sensitivity factors provided by the spectrometer manufacturer PHI [22].
The transmission FTIR spectra were obtained with a
Nicolet 20SXB spectrophotometer using KBr wafers containing about 0.5 g carbon. These wafers were dried
overnight at 393 K before the spectra were recorded. The
FTIR spectra of the samples were obtained by adding 100
scans at a resolution of 1 cm 21 .
Both total surface acidity and basicity were determined
by titration with NaOH and HCl, respectively, following
the method described elsewhere [8]. The pH PZC of the
activated carbons [6] was obtained by mixing 1 g of
carbon with 20 cm 3 of CO 2 -free distilled water. The slurry

was kept in a plastic bottle and shaken periodically for one


or two days until the pH had stabilized, taking the final pH
of the slurry as the pH PZC of the solid. Finally, the surface
acidity was also evaluated by studying the dehydration
reaction of methanol at 453 K. Catalytic test was carried
out in a glass microreactor at atmospheric pressure and the
catalyst was heat treated at 453 K for 2 h, in He flow,
before studying its activity. The reaction was performed in
He flow saturated with alcohol at 273 K. The total flow
was 63 cm 3 min 21 and the space velocity 0.41 h 21 .
Analysis of reaction products was done by on-line gas
chromatography using a Perkin-Elmer gas chromatograph,
model 8500, with flame ionization detector and a column
Carbopack B80 / 120.

3. Results and discussion


Surface area and porosity of the original and oxidized
activated carbons are shown in Table 1. SN 2 of the original
samples increased with the degree of activation given to
the carbonized raw material. The oxidation treatments did
not change SN 2 in the case of BV5. However, in the case of
BV15 and BV29 there was a decrease in this parameter
with oxidation treatment with greater degrees of activation
of the original carbon resulting in a more pronounced
decrease in SN 2 . This occurs because an increase in the
degree of activation makes the pore walls thinner and,
thus, more easily destroyed by the oxidizing agent [7]. V2
value decreased significantly with oxidation treatment in
all cases, except in the case of BV29H. It is noteworthy
that whereas the S and N treatments decreased the macropore volume, V3 , of the original sample, H treatment
increased the macropore volume.
Table 1
Surface area and porosity of the original and oxidized activated
carbons
Sample

SN 2
m 2 g 21

V2 a
cm 3 g 21

V3 b
cm 3 g 21

BV5
BV5H
BV5S

467
425
497

0.17
0.16
0.16

0.18
0.22
0.13

BV15
BV15H
BV15S

716
651
646

0.19
0.17
0.17

0.29
0.42
0.36

BV29
BV29H
BV29S
BV29N

928
769
780
738

0.14
0.20
0.13
0.15

0.74
0.83
0.64
0.59

a
V2 is the pore volume of pores with diameter between 3.6 and
50 nm.
b
V3 is the pore volume of pores with diameter greater than 50
nm.

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001

1997

Table 2
Amounts of CO and CO 2 evolved after heating up to 1273 K in
He flow. Total and surface oxygen concentrations
[O] T a
%

[O] S b
%

Sample

CO
mmol g 21

CO 2
mmol g 21

CO / CO 2

BV5
BV5H
BV5S

0.45
1.15
0.90

0.11
0.35
0.23

4.09
3.29
3.91

1.1
3.0
2.2

BV15
BV15H
BV15S

0.40
1.97
2.16

0.10
0.60
0.45

4.00
3.28
4.80

1.0
5.1
4.9

6.7
10.2
8.7

BV29
BV29H
BV29S
BV29N

0.37
2.90
2.87
5.28

0.09
0.90
0.61
2.22

4.11
3.22
4.70
2.38

0.9
7.5
6.5
15.6

4.6
12.0
9.1
12.7

a
Total oxygen content from the amounts of CO and CO 2
evolved.
b
Surface oxygen content from XPS measurements.

From TPD spectra, the amounts of CO and CO 2 evolved


up to 1273 K were obtained and these are compiled in
Table 2. The results show that the S treatment fixed the
lowest amount of CO 2 -evolving surface complexes, i.e.
carboxyl and lactone groups, whereas the N treatment fixed
the largest amount of these groups. Therefore, the CO /
CO 2 ratio decreases in the order S.H.N. From the
amounts of CO and CO 2 evolved, the total oxygen content,
[O] T , was obtained. The N treatment fixed the largest
amount of oxygen, whereas the S treatment fixed the
smallest amount.
The surface oxygen content, [O] S , was obtained from
XPS. Papirer et al. [23] indicate that, with a first approximation, the 8090% peak area of the O 1s peak originates
from the contribution of surface atoms, or atoms of the
3.54 nm thick outer shell and the C 1s peak from C atoms
located at the surface or in the 45 nm outer shell. Results
found show that for the original carbons, BV15 and BV29,
the oxygen groups were essentially concentrated on the
external surface of the carbon particles. When these
carbons were oxidized with both H 2 O 2 and (NH 4 )S 2 O 8
most of the oxygen surface complexes were fixed on the
external surface and so [O] S .[O] T . However, the N
treatment gave a [O] S ,[O] T and, therefore, was able to fix
more oxygen surface complexes on the internal surface of
the carbon particles than the other two treatments.
The chemical nature of the oxygen surface complexes
fixed on the samples after the oxidizing treatments was
studied both by XPS and FTIR of selected samples. XPS
gives information about the composition of the most
external surface of the activated carbon particles, as
mentioned above. Fig. 1 shows, as an example, the fits of
C 1s of sample BV29H. With all carbons, the C 1s signal
showed an asymmetric tailing. This is partially due to the
intrinsic asymmetry of the graphite peak [24,25] and to the
contribution of oxygen surface complexes. For the curve

Fig. 1. XPS pattern (C 1s core level) of sample BV29H.

fitting of the measured peaks, the following functions were


considered: aromatic and aliphatic carbon, single CO
bonds, double C=O bonds, carboxyl groups and carbonate.
The contribution of plasmon was negligible. Table 3 shows
the binding energy (BE) values adopted in C 1s fits [25
28].
The relevant peak areas are compiled in Table 4. Listed
values were normalized to the area of peak 1 arbitrarily set
to 100. Results show that, on the most external surface of
the samples, the S treatment fixed the lowest amount of
double C=O bonds and carboxyl groups, peaks 3 and 4,
respectively. This lowest amount of carboxyl groups was
also detected by TPD (Table 2). Thus, the concentration of
carboxyl groups was lowest after the S treatment both on
the total and the most external surface of the particles. The
S treatment also yielded the lowest surface concentration
of groups with a single CO bond. The surface concentration of carbonate is almost the same after the three
oxidizing treatments. Therefore, these results clearly indicate that the surface concentration of acid groups such as
carboxyl, lactone or phenol groups was the lowest after the
S treatment.
FTIR of sample BV29 and its oxidized derivatives are
depicted in Fig. 2, as an example. Bands in the region of
18601650 cm 21 are attributed to the C=O stretching
vibrations corresponding to carbonyl and carboxyl groups
[2931]. In this region, the oxidized carbons present more
pronounced bands than the original carbon. Samples
BV29S and BV29N show a band at 1740 cm 21 which can
be assigned to lactone groups [2932], whereas in sample
BV29H this band is shifted to 1720 cm 21 . This band could
be assigned either to lactone groups [2931] or to non-

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001

1998
Table 3
Position and assignments of C 1s peaks
Peak number

Binding energy
(eV)

Assignment

1
2

284.6 285.1
286.1

287.3

4
5a
6

289.1
290.6
291.6

Aromatics and aliphatics


Single CO bond (alcohol, ether, phenol,
COH of an enolketo group)
Double bonded carbon oxygen groups
(carbonyl, quinone)
Carboxyl groups
Carbonate, CO 2
Plasmon

This peak is also attributed [23] to shake-up satellite (pp*).

aromatic carboxyl groups, for which the C=O stretching


has been reported [3033] to occur at 1712 cm 21 .
Sample BV29H shows a shoulder at 1700 cm 21 and a
peak at 1660 cm 21 , sample BV29N only shows a well
defined peak at 1660 cm 21 and sample BV29S shows a
doublet with peaks at 1680 and 1660 cm 21 . It has been
reported [33,34] that the C=O stretching vibration of
carboxyl acid groups existing in an aromatic ring structure
generally appear at 17001680 cm 21 and its wave number
is affected by different peripheral groups. Finally, the band
at 1660 cm 21 can be attributed to quinone or conjugated
ketone [9,29,33,35].
The band at 1580 cm 21 , that appears in the original and
oxidized samples, has been observed by many authors and
has not been interpreted unequivocally. This has been
assigned to aromatic ring stretching coupled to highly
conjugated carbonyl groups (C=O) [29]. The 1460 cm 21
can be attributed either to an OH deformation vibration in
carboxyl groups or CH bending vibrations [9,29,36], and
appears in all oxidized samples. The bands in the region of
14001380 cm 21 , that also appear in oxidized samples,
can be ascribed either to carboxylcarbonate structures
[29,36] or to aromatic C=C bond and various substitution
modes of the aromatic ring [34].
Bands in the 12001000 cm 21 region are difficult to
assign because here a number of broad overlapping bands

are superimposed. They can not, therefore, be described in


terms of simple motion of specific functional groups or
chemical bonds [30]. At 1040 cm 21 , sample BV29H
presents a shoulder and samples BV29N ad BV29S a
defined peak which corresponds to alcoholic CO stretch-

Table 4
Comparison of C 1s peak areas normalized to the area of peak 1
arbitrarily set to 100
Sample

C 1s peak number
1

BV15
BV15H
BV15S

100
100
100

1.94
7.55
5.69

1.31
3.72
1.92

1.69
3.13
1.69

2.60
2.46
3.01

BV29
BV29H
BV29S
BV29N

100
100
100
100

1.38
3.74
3.44
7.82

0.88
4.65
1.80
3.87

1.26
2.83
1.88
2.69

1.60
2.61
2.67
2.65

Fig. 2. FTIR spectra for activated carbons from BV29 series.

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001

ing vibration [33]. Bands below 950 cm 21 are characteristic of out-of-plane deformation vibrations of CH groups
in aromatic structures [29,3638].
In conclusion, qualitatively, the FTIR technique detected
lactone, quinone or conjugated quinone, carboxylcarbonate structures and alcohol groups in all oxidized samples.
In the case of samples obtained after the H and S
treatments a band corresponding to carboxyl groups
bonded to aromatic structures also appears.
The acidbase character of samples is shown in Table 5.
Total surface acidity increased with the oxidation treatment, whereas the total surface basicity decreased. This
phenomenon has been observed before and is considered,
by different authors [1119], as an indication that the basic
surface sites are of Lewis type and associated with p
electron-rich regions within the basal planes. The greatest
total surface acidity was obtained after the N treatment,
because this fixed the largest amount of oxygen surface
complexes. However, it is important to note that after the S
treatment the lowest total surface basicity was obtained, in
spite of the fact that this treatment fixed the lowest amount
of oxygen, as shown above.
It is noteworthy that the S treatment always yielded
activated carbons with the lowest pH PZC . This result has
been reported before with other activated carbons obtained
from almond shells [7,8] and indicates that the S treatment
fixed the oxygen surface complexes with the highest acid
strength, in spite of the fact that this treatment fixed the
lowest amount of oxygen surface complexes that evolved
as CO 2 and the lowest [O] T .
The acidbase character of the carbon surface can also
be determined by indirect methods, such as the measurement of their catalytic activity in the decomposition of
alcohols as proposed for metal oxides and other inorganic
compounds [3944]. In this case we have used the
decomposition of methanol to test the acid character of
selected samples. The decomposition reaction of alcohols
catalyzed by activated carbons has been studied by differ-

Table 5
Total surface acidity and basicity and pH PZC of the original
activated carbons and their oxidized derivatives
Sample

Total surface acidity


meq NaOH g 21

Total surface basicity


meq HCl g 21

pH PZC

BV5
BV5H
BV5S

0.04
0.09
0.13

0.30
0.28
0.26

9.97
6.50
3.09

BV15
0.02
BV15H 0.59
BV15S 0.70

0.45
0.37
0.25

BV29
BV29H
BV29S
BV29N

0.69
0.48
0.15
0.24

0.00
0.96
0.81
2.19

1999

ent authors [10,4548], and one of the conclusions reached


was that the dehydration reaction took place on acid sites
of carboxyl type essentially placed at the external surface
of the carbon particles.
The only product obtained in the decomposition of
methanol by the activated carbons studied was dimethyl
ether (DME) and results found with selected catalysts at a
reaction temperature of 453 K are shown in Table 6. The
original activated carbons, BV15 and BV29, had no
activity to produce DME due to the basic character of their
surface. However, for the oxidized activated carbons the
conversion and activity to produce DME increased when
the pH PZC of the carbon decreased. Thus, the highest
activity was obtained after the S treatment. These results
clearly show again that the highest acid strength was
obtained after this treatment.
It is well known [49], that the acid strength of carboxylic acids is affected by the nature of the groups which
are close neighbors of the carboxyl carbon. This phenomenon is commonly called the inductive effect, and it is
taken as negative effect if the substituent is acid-enhancing, and positive if the substituent is acid-weakening. The
negative inductive effect can be due either to the electronegativity or to a field effect of the a-substituent, in
which the dipole of the a-substituent produces an electrostatic field at the carboxylic proton, which favors ionization. In general, groups that withdraw electrons increase
acidity and decrease basicity, while electron-donating
groups act in the opposite direction. As would be expected
the influence falls off with increasing distance of the
substituent from the carboxyl group. Many substituents
and groups exhibit an acid-enhancing effect and among the
fuctionalities that can be found on the surface of the
activated carbons are carbonyl and hydroxyl groups. Thus,
the pK of acetic, hydroxyacetic and oxoacetic acids [50]
are 4.75, 3.83 and 3.30, respectively. In the case of
aromatic acids [51] resonance effects can be also important, and in general, they lead to the same results as the
inductive effect. That is, here too, electron-withdrawing
groups increase acidity and decrease basicity. As a result
of both, inductive and resonance effects, charge dispersal
leads to a greater stability of the carboxylate anion. Thus,
Table 6
Conversion and activity, r DME , of the activated carbons in the
dehydration reaction of methanol at 453 K
Sample

Conversion
%

r DME
mmol g 21 min 21

9.90
4.70
2.55

BV15
BV15H
BV15S

nil
0.03
4.48

nil
0.06
9.69

9.80
4.11
2.72
3.79

BV29
BV29H
BV29S
BV29N

nil
0.04
2.82
0.65

nil
0.10
6.00
1.40

2000

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001

the pK of benzoic and o-hydroxybenzoic acids [51] are


4.19 and 2.97, respectively.
One can speculate that the highest acid strength and
lowest total surface basicity of the activated carbons after
the S treatment could be due to this treatment favors the
fixation of carboxyl groups adjacent to the above acidenhancing groups. This would be possible with this
treatment because, as has been shown, of the oxidizing
treatments studied this one fixed the smallest amount of
surface carboxyl groups and there would, therefore, be a
greater probability of finding carboxyl groups adjacent to
other oxygen surface complexes that enhance their acidity.

4. Conclusions
This study shows the changes in surface chemistry of
several activated carbons oxidized by H 2 O 2 , (NH 4 ) 2 S 2 O 8
and HNO 3 . The highest total oxygen content was obtained
after the N treatment and the lowest one after the S
treatment. Both H and S treatments fixed most of the
oxygen groups on the external surface of the carbon
particles and the N treatment on the internal surface.
The techniques XPS and FTIR detected oxygen surface
groups with single CO bond, lactone, carboxyl, quinone
or conjugated ketone and carboxylcarbonate structures in
all oxidized samples. Oxidized samples obtained after the
(NH 4 ) 2 S 2 O 8 treatment presented the lowest surface concentration in acidic functional groups such as: carboxyl,
lactone and phenolic groups.
The highest total surface acidity was obtained after the
N treatment and the lowest surface basicity after the S
treatment. However, the highest acid strength was obtained
after the S treatment, as was clearly indicated by the
measurements of both pH PZC and catalytic activity in the
dehydration reaction of methanol, in spite of the fact that
this treatment fixed the lowest amount of both oxygen and
surface acid groups. The highest acid strength and lowest
surface basicity could be due to the fact that this treatment
favors the fixation of carboxyl groups close to other groups
that, due to both negative inductive effects and resonance
effect, enhance their acidity.

Acknowledgements
The authors extend their gratitude to Doctor E. Rod
from Dpto. Qumica

rguez-Castellon
Inorganica,
Uni
versidad de Malaga,
for the XPS measurements. This work

General de Ensenanza
was supported by the Direccion
Cientfica

Superior e Investigacion
Project No. PB97-0831.

References
[1] Wen WW, Sun SC. An electrokinetic study on the oil
flotation
of
oxidized
coal.
Sep
Sci
Technol
1981;16(10):1491521.

[2] Lau AC, Furlong DN, Healey TW, Grieser F. The electrokinetic properties of carbon black and graphitized carbon
black aqueous colloids. Colloids Surf 1986;18(1):93104.
[3] Kinoshita K. In: Carbon: electrochemical and physicochemical properties, New York: Wiley, 1988, pp. 86166.
[4] Fuerstenau DW, Rosenbaum JM, You YS. Electrokinetic
behavior of coal. Energy Fuels 1988;2(3):2415.
y Leon
CA, Osseo-Asare K, Radovic LR. On
[5] Solar JM, Leon
the importance of the electrokinetic properties of carbons for
their use as catalyst supports. Carbon 1990;28(23):36975.
y Leon
CA, Radovic LR. Interfacial chemistry and
[6] Leon
electrochemistry of carbon surfaces. In: Thrower PA, editor,
Chemistry and physics of carbon, vol. 24, New York:
Dekker, 1994, pp. 213310.
MA, Joly JP, Bautista[7] Moreno-Castilla C, Ferro-Garca
F, Rivera-Utrilla J. Activated
Toledo I, Carrasco-Marn
carbon surface modifications by nitric acid, hydrogen peroxide, and ammonium peroxydisulfate treatments. Langmuir
1995;11(11):438692.
F, Mueden A. The
[8] Moreno-Castilla C, Carrasco-Marn
creation of acid carbon surfaces by treatment with
(NH 4 ) 2 S 2 O 8 . Carbon 1997;35(1011):161926.
F, Maldonado-Hodar

[9] Moreno-Castilla C, Carrasco-Marn


FJ,
Rivera-Utrilla J. Effects of non-oxidant and oxidant acid
treatments on the surface properties of an activated carbon
with very low ash content. Carbon 1998;36(12):14551.
F, Mueden A, Moreno-Castilla C. Surface[10] Carrasco-Marn
treated activated carbons as catalysts for the dehydration and
dehydrogenation reactions of ethanol. J Phys Chem B
1998;102(6):923944.

MV, Stoeckli F, Moreno-Castilla C, Carrasco[11] Lopez-Ramon


F. On the characterization of acidic and basic surface
Marn
sites on carbons by various techniques. Carbon
1999;37(8):121521.
[12] Puri BR. Surface complexes on carbons. In: Walker PL,
editor, Chemistry and physics of carbon, vol. 6, New York:
Dekker, 1970, pp. 191282.
[13] Studebaker ML. Chemistry of carbon black and reinforcement. Rubber Chem Technol 1957;39:140083.
[14] Fabish TJ, Schleifer DE. Surface chemistry and the carbon
black work function. Carbon 1984;22(1):1938.
y Leon
CA, Solar JM, Calemma V, Radovic LR.
[15] Leon
Evidence for the protonation of basal plane sites on carbon.
Carbon 1992;30(5):797811.

[16] Menendez
JA, Phillips J, Xia B, Radovic LR. On the
modification and characterization of chemical surface properties of activated carbon: In the search of carbons with
stable basic properties. Langmuir 1996;12(18):440410.
[17] Barton SS, Evans MJB, Halliop E, MacDonald JAF. Acidic
and basic sites on the surface of porous carbon. Carbon
1997;35(9):13616.
MA, Menendez

[18] Montes-Moran
JA, Fuente E, Suarez
D.
Contribution of the basal planes to carbon basicity: An Ab
initio study of the H 3 O 1 p interaction in cluster models. J
Phys Chem B 1998;102(29):5595601.

MV, Stoeckli F, Moreno-Castilla C, Carrasco[19] Lopez-Ramon


Mar F. Specific and non-specific interactions of water
molecules with carbon surfaces from immersion calorimetry.
Carbon, in press.
MA, Utrera-Hidalgo E, Rivera-Utrilla J,
[20] Ferro-Garca
Moreno-Castilla C, Joly JP. Regeneration of activated carbons exhausted with chlorophenols. Carbon 1993;31(6):857
63.

C. Moreno-Castilla et al. / Carbon 38 (2000) 1995 2001


[21] Shirley DA. High-resolution X-ray photoemission spectrum
of the valence bands of gold. Phys Rev B 1972;5(12):4709
14.
[22] Physical Electronics, 6509 Flying Cloud Drive, Eden Prairie,
Mn 55344, USA.
[23] Papirer E, Lacroix R, Donnet JB, Nanse G, Fioux P. XPS
Study of the halogenation of carbon black Part 1.
Bromination. Carbon 1994;32(7):134158.
[24] Cheung TTP. X-ray photoemission of polynuclear aromatic
carbon. J Appl Phys 1984;55(5):138893.
[25] Darmstadt H, Roy C, Kaliaguine S. ESCA Characterization
of commercial carbon blacks and of carbon blacks from
vacuum pyrolysis of used tires. Carbon 1994;32(8):1399
406.
[26] Desimoni E, Casella GI, Salvi AM. XPS / XAES study of
carbon fibers during thermal annealing under UHV conditions. Carbon 1992;30(4):5216.
[27] Desimoni E, Casella GI, Salvi AM, Cataldi TRI, Morone A.
XPS Investigation of ultrahigh-vacuum storage effects on
carbon fiber surfaces. Carbon 1992;30(4):52731.

[28] Zielke U, Huttinger


KJ, Hoffman WP. Surface-oxidized
carbon fibers: I. Surface structure and chemistry. Carbon
1996;34(8):98398.
[29] Zawadzki J. Infrared spectroscopy in surface chemistry of
carbons. In: Thrower PA, editor, Chemistry and physics of
carbon, vol. 21, New York: Dekker, 1989, pp. 147380.
[30] Painter P, Starsinic M, Coleman M. Determination of functional groups in coal by Fourier transform interferometry. In:
Fourier transform infrared spectroscopy, vol. 4, New York:
Academic Press, 1985, pp. 16989.
[31] Fanning PE, Vannice MA. A DRIFTS study of the formation
of surface groups on carbon by oxidation. Carbon
1993;31(5):72130.
[32] Zhuang QL, Kyotani T, Tomita A. DRIFT and TK / TPD
analyses of surface oxygen complexes formed during carbon
gasification. Energy Fuels 1994;8:7148.
[33] Shin S, Jang J, Yoon SH, Mochida I. A study on the effect of
heat treatment on functional groups of pitch based activated
carbon fiber using FTIR. Carbon 1997;35(12):173943.
[34] Pretsch E, Clerc T, Seibl J, Simon W. In: Tables of spectral
data for structure determination of organic compounds, 2nd
ed, New York: Springer-Verlag, 1989, pp. I5I265.
[35] Starsinic M, Taylor RL, Walker Jr. PL, Painter PC. FTIR
Studies of Saran chars. Carbon 1983;21(1):6974.
[36] Biniak S, Szymanski G, Siedlewski J, Swiatkowski A. The
characterization of activated carbons with oxygen and nitrogen surface groups. Carbon 1997;35(12):1799810.

2001

[37] Meldrum BJ, Rochester CH. In situ infrared study of the


surface oxidation of activated carbon in oxygen and carbon
dioxide. J Chem Soc Faraday Trans 1990;86(5):8615.
[38] Meldrum BJ, Rochester CH. In situ infrared study of the
modification of the surface of activated carbon by ammonia,
water and hydrogen. J Chem Soc Faraday Trans
1990;86(10):18814.
[39] Ai M. Oxidation activity and acidbase properties of tin
dioxide-based binary catalysts. I. Tin dioxidevanadium
pentoxide system. J Catal 1975;40(3):31826.
[40] Ai M. Oxidation activity and acidbasic properties of tin
dioxide-based binary catalysts. II. Tin dioxidemolibdenum
trioxide and tin dioxidephosphorus pentoxide systems. J
Catal 1975;40(3):32733.
[41] Gervasini A, Auroux A. Acidity and basicity of metal oxide
surfaces. J Catal 1991;131:1908.
[42] Afanasiev P, Geatent C, Breysse M, Coudurier G, Vedrine JC.
Influence of preparation method on the acidity of
MoO 3 (WO 3 ) / ZrO 2 catalysts. J Chem Soc Faraday Trans
1994;90(1):193202.
[43] Gervasini A, Bellussi G, Fenyvesi J, Auroux A. Microcalorimetric and catalytic studies of the acidic character
of modified metal oxide surfaces. 1. Doping ions on alumina,
magnesia and silica. J Phys Chem 1995;99(14):511725.

[44] Moreno-Castilla C, Maldonado-Hodar


FJ, Rivera-Utrilla J,

Rodrguez-Castellon
E. Group 6 metal oxidecarbon
aerogels. Their synthesis characterization and catalytic activity in the skeletal isomerization of 1-butene. Appl Catal A
1999;183:34556.
[45] Szymanski GS, Rychlicki G. Importance of oxygen surface
groups in catalytic dehydration and dehydrogenation of
butan-2-ol promoted by carbon catalysts. Carbon 1991;29(4
5):48998.
[46] Grunewald GC, Drago RS. Carbon molecular sieves as
catalysts and catalyst supports. J Am Chem Soc
1991;113(5):16369.
[47] Szymanski GS, Rychlicki G. Catalytic conversion of propan2-ol on carbon catalysts. Carbon 1993;31(2):24757.
[48] Szymanski GS, Rychlicki G, Terzyk AP. Catalytic conversion of ethanol on carbon catalysts. Carbon 1994;32(2):265
71.
[49] Roberts JD, Caserio MC. In: Basic principles of organic
chemistry, New York: Benjamin Inc, 1965, pp. 50764.
[50] Lide RD, editor, Handbook of chemistry and physics, 72nd
ed, Boca Raton: CRC Press, 1992, pp. 8.3940.
[51] March J. In: Advanced organic chemistry, 4th ed, New York:
Wiley, 1992, pp. 2639.

Вам также может понравиться