Sie sind auf Seite 1von 6

Catalysis Today 218219 (2013) 5156

Contents lists available at ScienceDirect


Catalysis Today
j our nal homepage: www. el sevi er . com/ l ocat e/ cat t od
Carbon as a catalyst: Esterication of acetic acid with ethanol
Raquel P. Rocha, Manuel F.R. Pereira, Jos L. Figueiredo

LCM Laboratory of Catalysis and Materials Associate Laboratory LSRE/LCM, Faculdade de Engenharia, Universidade do Porto, Rua Dr. Roberto Frias,
4200-465 Porto, Portugal
a r t i c l e i n f o
Article history:
Received 14 December 2012
Received in revised form27 March 2013
Accepted 29 September 2013
Available online 25 October 2013
Keywords:
Carbon nanotubes
Carbon xerogels
Sulphonic acid groups
Carbon catalyst
Esterication
Acetic acid
a b s t r a c t
Multiwalled carbon nanotubes (CNTs) and carbon xerogels prepared by the solgel process (CXs) were
modied by different chemical and thermal treatments with sulphuric and nitric acid in order to produce
materials with different textural properties and acidic nature. The presence of sulphonic acid groups was
identied by X-ray photoelectron spectroscopy on the sulphuric acid treated samples. Carboxylic acids
and anhydrides, phenol groups, some lactones and carbonyl-quinone groups were incorporated during
the nitric acid treatment, as shown by temperature programmed desorption. Oxidation with nitric acid
affects more the textural properties (the effect is more pronounced in the case of CNTs than in CXs)
promoting an increase of surface area and a decrease of total pore volume, while a decrease of surface area
was observed with the sulphuric acid treatment. The prepared materials were tested in the esterication
of acetic acid with ethanol as a model acid-catalyzed reaction. The rate of formation of ethyl acetate was
found to depend strongly on the chemical properties of the carbon materials, the presence of sulphonic
acid groups being more adequate for this reaction. A good correlation was obtained between the rate of
formation of ethyl acetate and the concentration of sulphonic groups.
2013 Elsevier B.V. All rights reserved.
1. Introduction
The versatility of carbon as a catalyst was recognized long
time ago [1]. In fact, carbon materials can perform well in
reactions which are typically catalyzed by metals (such as dehy-
drogenations), non-stoichiometric oxides (such as oxidations and
reductions) and acids (such as alkylation and dehydration) as dis-
cussed in a recent review [2]. The performance of any catalyst
depends on the availability of suitable active sites, capable of
chemisorbing the reactants and forming surface intermediates of
adequate strength. Therefore, the catalytic properties of carbon
materials are mostly determined by their surface chemistry [3].
The presence of heteroatoms (O, N, H, Cl, S, etc.) bound to the edges
of the graphene layers in the carbon material originates a variety
of surface functional groups. The concentration of these groups can
be further tuned in order to suit specic requirements [2,4]. Acidic
groups include carboxylic acids and anhydrides, lactones or lactols,
and phenols, and can be easily incorporated by oxidative treat-
ments with nitric acid, oxygen peroxide and oxygen [58]. Carbon
materials can also be functionalized with sulphonic acid groups,
providing efcient acid catalysts [913].
Suitable methods of analysis of the carbon functional groups
are now available, including TPD-MS and XPS [4,14]. Therefore, it

Corresponding author. Tel.: +351 225081663.


E-mail address: jlg@fe.up.pt (J.L. Figueiredo).
shouldbe possible tocorrelate the catalytic properties of the carbon
materials withtheir surfacechemistry. However, this is not thegen-
eral case, quantitative relationships being published only in a few
cases [16]. The present work explores the effect of different surface
chemistry of carbon nanotubes and carbon xerogels in the esteri-
cation of acetic acid, as a model acid-catalyzed reaction. Carbon
materials with sulphonic acid groups ( SO
3
H) have been reported
as excellent solid acid catalysts for this reaction [10,12,1517].
2. Experimental
2.1. Materials
Different chemical andthermal treatments wereappliedtomul-
tiwalled carbon nanotubes (CNTs) and carbon xerogels (CXs) in
order to produce materials with different textural and chemical
properties. The pristine CNTs (CNT-O) with an average diameter
of 9.5nm, average length of 1.5mand carbon purity higher than
95% were purchased from NANOCYL
TM
(NC3100 series). The CXs
were produced by polycondensation of resorcinol with formalde-
hyde, adaptingtheproceduredescribedelsewhere[18]. Briey, 25g
of resorcinol were added to 40mL of distilled water under stir-
ring and 34mL of formaldehyde were introduced after complete
dissolution. The pHwas adjustedto6.0(inorder toproduce a meso-
porous structure [18]) by dropwise addition of a sodiumhydroxide
solution. The gelation step was completed after 3 days at 85

C in a
parafn bath. Once nished, the material was dried in an oven from
0920-5861/$ see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cattod.2013.09.049
52 R.P. Rocha et al. / Catalysis Today 218219 (2013) 5156
Table 1
Textural properties of the original carbon nanotube and xerogel samples and selected modied samples.
Sample SBET (m
2
g
1
) Vp (cm
3
g
1
) SMeso (m
2
g
1
) V
Micro
(cm
3
g
1
) dp (nm)
CNT-O 302 2.90
CNT-N7 358 1.62
CNT-N10 330 1.32
CNT-Nc 312 0.54
CX-O 672 1.16 244 0.173 17.4
CX-N7 697 1.14 234 0.188 17.5
CX-S1 416 0.95 213 0.084 17.3
CX-S2 528 1.04 222 0.125 17.4
60 to 120

C during 4 days, increasing the temperature 20

C/day.
After drying, the material was further subjected to a carbonization
procedure: using a heating ramp of 2

Cmin
1
under nitrogen ow
(100cm
3
min
1
), the material was heated until 150, 400, 600 and
800

C holding each temperature between 1 and 6h (higher tem-


perature), resulting ina carbonxerogel henceforwardreferredto as
CX-O. Nitric acidoxidationandsulphuric acidtreatments were per-
formed in order to introduce different types and amounts of acidic
functional groups on the surface of the distinct carbon structures.
In a typical liquid-phase oxidation, CNT-O was treated with nitric
acid (ratio 1:75 (w/v)) at boiling temperature during 3h, using dif-
ferent acidconcentrations (7, 10Mandconcentrated). The oxidized
material was washed several times with distilled water until neu-
tral pHwas reachedandnally driedat 110

Cinanovenovernight.
The resulting samples were designated in terms of the nitric acid
concentration as CNT-N7, CNT-N10 and CNT-Nc. The same proce-
dure was repeated using the CX-O sample; however, in this case,
only the 7M solution of nitric acid was tested (CX-N7 sample).
The liquid-phase chemical treatment with concentrated sulphuric
acid (H
2
SO
4
) was performed in a round-bottomask immersed in
a parafn bath at 150

C during 6h. Using the CX-O sample two


treatments were done using different ratios of catalyst weight/acid
volume: 1g:150mL H
2
SO
4
(CX-S1) and 1g:20mL H
2
SO
4
(CX-S2).
The samples were washed withdistilled water until neutral pHwas
reached, and nally dried at 110

C in an oven during the night.


Lastly, one portion of the CX-S1 sample was subjected to a nal
thermal treatment up to 250

C using helium as inert gas (heated


at 10

Cmin
1
; He ow of 25cm
3
min
1
) with the aim of remov-
ing part of the surface groups incorporated during the sulphuric
treatment, sample CX-S1-250. Representative samples of CNTs and
CXs were subjected to a thorough characterization and tested in
the catalytic esterication of acetic acid with ethanol.
2.2. Catalyst characterization
The textural characterization of the carbon materials was based
on the nitrogen adsorption-desorption isotherms. The samples
were pre-degassed at 150

C for 5h and the analysis performed in


a Quantachrome NOVA 4200e multi-station apparatus at 196

C.
Surface areas of the carbon samples were determined according
to the Brunauer, Emmett and Teller (BET) method and the aver-
age mesoporous diameters (d
p
) of the CX samples were obtained
by the method of Barrett, Joyner and Halenda (BJH). The mesopore
surface area (S
meso
), andmicropore volume (V
micro
) were calculated
using t-method, as described elsewhere [4]. The surface chemistry
of the samples was characterized by temperature programmed
desorption (TPD) analysis in a fully automated AMI-300 Catalyst
Characterization apparatus (Altamira Instruments) connected to a
Dycor Dymaxion Mass Spectrometer. The samples (0.100g) were
placed in a U-shaped quartz tube located inside an electrical fur-
nace and heated up to 1100

C at 5

Cmin
1
using a constant ow
rate of helium equal to 25cm
3
min
1
. The CO and CO
2
released
during TPD were monitored, and calibration with the respective
gases was carried out at the end of each analysis. SO and SO
2
signals were also followed in the case of sulphuric-treated sam-
ples. Elemental analysis was performed in a Carlo Erba EA 1108
Elemental Analyser. X-ray photoelectron spectroscopy (XPS) anal-
ysis was performed in a VG Scientic ESCALAB 200A spectrometer
using a non-monochromatized Mg K radiation (1253.6eV), the
binding energies being calibrated with the respect to the C1s peak
at 285.0eV. The characterizationwas completedwiththermogravi-
metric analysis (TGA), by heating the samples at 10

Cmin
1
under
nitrogen ow (50cm
3
min
1
) up to 900

C using a STA 490 PC/4/H


Luxx Netzsch thermal analyzer.
2.3. Experimental procedure
The catalytic activity of the prepared materials was tested on
the esterication of acetic acid with ethanol. The experiments were
carried in a round-bottom ask with necks for sample extraction
and for coupling a condenser, immersed in a parafn bath. 0.1mol
of acetic acid, 1mol of ethanol and 0.20g of catalyst were placed
inside the reactor and nally the temperature was set at 70

C. The
carbon material was recovered at the end of each run for further
characterization, or for reutilization in cyclic experiments. A blank
experiment without catalyst was also performed in the same con-
ditions. For comparison, homogeneous catalysis using H
2
SO
4
as
catalyst was also studied.
2.4. Analytical techniques
Samples periodicallywithdrawnfromthereactor wereanalyzed
by high performance liquid chromatography (HPLC) with a Hitachi
Elite LaChrom system equipped with ultraviolet (=210nm) and
refractive index detectors in series. A Bio-Rad Aminex HPX-87H
column (300mm7.8mm) and a 4mM sulphuric acid solution as
mobile phase (owrate of 0.8mL min
1
) were used.
3. Results and discussion
3.1. Carbon materials characterization
Table 1 summarizes the textural properties of the original and
modied CNT and CX samples determined from the N
2
adsorp-
tion isotherms. In the case of the CNT samples treated with nitric
acid, the differences between the S
BET
surface areas are smaller
than 100m
2
g
1
. However, the total pore volume (V
p
, determined
fromthe N
2
uptake at p/p
0
=0.99) decreases byincreasingthe HNO
3
concentration.
According to IUPAC, the isotherms of the CNT samples (Fig. 1a)
can be classied as type II, characteristic of nonporous materi-
als such as CNTs. On this carbon material, the pore volume is
mostly associated to the free space occurring in CNT bundles [19],
therefore, the decrease of the pore volume could be related to the
combinedeffect of de-agglomerationof CNTs, promotedby the oxi-
dation treatment [20,21], and the incorporation of oxygen surface
groups. On the other hand, the changes in S
BET
may be connected
R.P. Rocha et al. / Catalysis Today 218219 (2013) 5156 53
0.0 0.5 1.0
0
400
800
1200
1600
2000
0.0 0.5 1.00.0 0.5 1.0
CNT-N10 CNT-O CNT-N7
V
a
d
s

(
c
m
3

g
-
1
,

S
T
P
)
a.
p/p
0
0.0 0.5 1.0
0
400
800
1200
0.0 0.5 1.00.0 0.5 1.0
p/p
0
V
a
d
s

(
c
m
3

g
-
1
,

S
T
P
)
b.
CX-S1 CX-N7 CX-O
Fig. 1. N
2
adsorptiondesorption isotherms at 196

C for original and selected


modied carbon samples: (a) CNT samples; (b) CX samples.
to changes in length and perfection of the nanotubes, as a result of
the applied treatments [22] or by opening the tips of the tubes [23].
Regarding the CX samples, they presented type IV isotherms
(Fig. 1b), characteristic of mesoporous solids, with average pore
diameter around 20nm (determined by BJH method). Likewise,
the BET surface of the sample treated with nitric acid increased,
in comparison with the original carbon xerogel, while it decreased
in the case of samples treated with sulphuric acid. On the latter
samples (CX-S1 and CX-S2), both the surface area and the micro-
pore volume decrease, indicating that the sulphuric acid treatment
contributes to reduce the microporosity of the surface, contrary to
what happens with nitric acid [24].
Elemental and XPS analyses were done for selected samples,
mainly in the case of sulphuric acid treated samples, in order to
identifytheamount andthenatureof thesurfacegroups introduced
during the treatment. Table 2 shows the results obtained.
Table 2
Elemental and XPS analyses of original and sulphuric acid treated CX samples.
Sample Elemental analysis XPS
%C %N %H %S %O
a
%C %N %O %S
CX-O 92.4 0.75 0.39 7.23 n.d.
CX-S1 74.7 0.16 1.26 2.10 22.0 79.9% 0.71% 16.2% 3.18%
CX-S2 75.5 0.08 0.89 1.59 22.0 81.0% 0.55% 16.6% 1.89%
CX-S1-250 74.6 0.14 0.28 0.17 24.8 89.4% 0.25% 9.7% 0.60%
a
Oxygen determined by difference; n.d. not determined.
160 164 168 172 176
a.
CX-S1
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
Binding Energy (eV)
160 164 168 172 176
b.
CX-S2
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
Binding Energy (eV)
160 164 168 172 176
CX-S1-250
I
n
t
e
n
s
i
t
y

(
a
.
u
.
)
Binding Energy (eV)
c.
Fig. 2. S2p XPS spectra for the sulphuric acid treated carbon xerogel samples.
The CX-S1 sample presented the highest sulphur content
(3.18%), as expected, since it was prepared with a higher ratio of
sulphuric acid (1:150), contrary to the CX-S2 sample that has 1.89%
(ratio of 1:20). The thermally treatedsample, CX-S1-250, presents a
residual amount of sulphur (only 0.60%) which shows that most of
the incorporated S-containing groups are released fromthe surface
of the carbon structure at lowtemperatures (below250

C).
The S2p spectra of CX-S1, CX-S2 and CX-S1-250 samples (Fig. 2)
show a common peak around 169eV which is the BE typically
attributed to sulphonic groups ( SO
3
H) [25,26]. The absence of
other peaks, especially at lower BE, allows to conclude that other
S-containinggroups (suchas the thiol groups, 163eV) are not incor-
porated during the treatment applied [27].
The amounts of CO and CO
2
released during the TPD analyses
(Table 3) indicate that the oxygen content incorporated during
sulphuric acid treatment is much lower than during nitric acid
oxidation, suggesting that the formation of oxygen-containing
functional groups is more favoured by nitric acid oxidation. This
is in agreement with what was demonstrated in a previous work
[28]. Considering samples CNT-N7 and CX-N7, larger amounts of
54 R.P. Rocha et al. / Catalysis Today 218219 (2013) 5156
Table 3
Concentration of the CO and CO
2
groups released during TPD of the carbon mate-
rials subjected to different liquid phase treatments compared with the amount of
volatiles released by TGA.
Sample CO (mol g
1
) CO
2
(mol g
1
) %Volatiles
a
CNT-O 299 105 2.3
CNT-N7 1453 1008 9.3
CNT-Nc 1592 1422 15.7
CX-S1 732 330 16.4
CX-S2 657 191 15.0
CX-N7 2414 1715 35.1
a
Determined by TGA in a dry ash free basis.
CO and CO
2
are released during TPD, indicating that carboxylic
acids and phenol groups, as well as some anhydrides, lactones and
carbonyl-quinones, are incorporated on the surface of the carbon
materials upon nitric acid treatment [2,4]. However, samples con-
tainingsulphonic groups (CX-S1andCX-S2) present lower amounts
of evolved CO and CO
2
, as shown in Table 3. The global amount
of groups incorporated during the distinct treatments can be esti-
mated by the amount of volatiles released during the TGA analyses
(Table 3). Those results show that, in spite of the smaller amounts
of CO and CO
2
releasing groups incorporated with the sulphuric
acid treatment, the sulphonic groups represent a large percentage
(by weight) of the total surface groups.
The MS signal of SO
2
(m/z =64) was followed during the TPD
analyses of the CX-S samples (Fig. 3c). The spectra reveal desorption
of S-containing groups from the surface in a specic temperature
range (200400

C) mainly due to the decomposition of sulphonic


groups intoSO
2
species [29]. Inagreement withtheprevious results
of XPS analyses, the amounts of SO
2
released decrease in the same
order as the S content, as follows: CX-S1, CX-S2 and CX-S1-250.
3.2. Catalytic activity for the esterication of acetic acid
The catalytic activity of the samples was evaluated in the ester-
ication of acetic acid with ethanol, as described above. Under the
operating conditions employed, the homogeneous non-catalytic
reaction can be neglected (Blank). Fig. 4a shows the ethyl acetate
formation during the reaction time using selected samples. The
sample functionalized with sulphuric acid (CX-S1) was the most
active, with a yield of 52% after 6h, followed by the CX-S2 and CNT-
Nc samples. For comparison purposes, the homogeneous catalysis
reaction using H
2
SO
4
was also performed using an amount of acid
density equivalent to that of the CX-S1 sample (1.80mmol g
1
,
determined by titration). However, it should be noted that the acid
density in the case of CX-S1 represents all acidic groups (including
carboxylic acids, that are measurable by the titration technique)
reason why the catalytic performance of the sample is lower than
the homogenous catalyst.
Since acidic groups of different nature were incorporated into
two different types of carbon structures, the samples were divided
in two groups for a better understanding of their catalytic per-
formance: S containing samples (S-samples) (CX-S1, CX-S2, and
CX-S1-250, in decreasing order of the S content) and nitric acid
functionalized samples (CNT-Nc, CX-N7, CNT-N10 and CNT-N7, in
decreasing order of the carboxylic acid groups, released as CO
2
).
Regarding the S-samples, a good correlation is obtained between
the rate of formation of ethyl acetate after 1h and the concentra-
tion of SO
3
H groups determined from XPS data (Fig. 4b). With
respect to the nitric treated samples, the lower rates observed can
be partiallyexplainedbythe lowacidityof the incorporatedgroups.
The catalytic activity relates well with the amount of carboxylic
acids determined by deconvolution of CO
2
spectra during the TPD
analysis, as also shown in Fig. 4b.
A brief comparison with published results shows that the
turnover frequency(TOF) of theCX-S1samplepreparedinthis work
(1.9min
1
) is competitive with similar catalytic performances
published in the literature, as single walled carbon nanotubes
(TOF =1.9min
1
) [12], and with conventional strong solid acid cat-
alysts (such as the protonated Naon resin NR50, TOF =1.5min
1
[15]). Thus, carbon xerogels with a strong acid character can be
used as catalysts for the esterication of acetic acid, and the cat-
alytic activity can be correlated with the amount of the strong
acid sites. This methodology is being extended to other types of
acid-catalyzed reactions.
The catalytic stability of the sample CX-S1 was studied in three
consecutive runs in order to evaluate the catalyst lifetime of the
900 800 700 600 500 400 300 200 100
0.0
0.4
0.8
1.2
CX-N7
a.
[
C
O
]

(

m
o
l

g
-
1

s
-
1
)
Temperature (C)
CX-S1
CNT-N7
900 800 700 600 500 400 300 200 100
0.0
0.4
0.8
1.2
b.
Temperature (C)
[
C
O
2
]

(

m
o
l

g
-
1

s
-
1
)
CX-S1
CNT-N7
CX-N7
900 800 700 600 500 400 300 200 100
0.00E+000
2.00E-008
4.00E-008
6.00E-008
S
O
2
,

M
S

S
i
g
n
a
l

(
t
o
r
r

g
-
1
)
c.
Temperature (C)
CX-S2
CX-S1
Fig. 3. TPD spectra of some CX and CNT samples treated with nitric or sulphuric acid: (a) CO release; (b) CO
2
release; (c) MS signal of groups released as SO
2
(m/z =64).
R.P. Rocha et al. / Catalysis Today 218219 (2013) 5156 55
0 400 800 1200
0.0
0.3
0.6
0.9
1.2
1.5
1.8
2.1
Sulphuric treated samples
Nitric treated samples
(Solid Symbols - CXs)
(Open Symbols - CNTs)
b.
a.
E
t
A
c

F
o
r
m
a
t
i
o
n

(
m
m
o
l

g
-
1

m
i
n
-
1
)
[SO
3
H] or [CarbAcid] (mol g
-1
)
0.0 0.5 1.0 1.5 2.0 2.5 3.0
6 4 2 0
0
25
50
75
100
E
t
h
y
l

A
c
e
t
a
t
e

(
m
m
o
l
)
Time (h)
1st Run
2nd Run
3rd Run
c.
1st Run
2nd Run
Ethyl Acetate Formation (mmol g
-1
min
-1
)
3rd Run
0 2 4 6
0
25
50
75
100
Blank
CX-N7
CNT-Nc
CX-S2
CX-S1
H
2
SO
4
E
t
h
y
l

A
c
e
t
a
t
e

(
m
m
o
l
)
Time (h)
Fig. 4. (a) Ethyl acetate formation during the reaction time (T =70

C; molar ratio:
ethanol/acetic acid=10; 0.2g of catalyst); (b) ethyl acetate formation rate after 1h
usingtreatedCXandCNTsamples as functionof therespectivesurfaceconcentration
of sulphonic/carboxylic acid groups; (c) rate of formation of ethyl acetate using the
CX-S1 sample in the cyclic experiments (inset: Ethyl acetate formation during the
reaction time).
sample, which is an important factor when a newcatalyst is being
tested. Fig. 4c shows the consecutive runs with fresh acetic acid
and with the materials recovered after each run, as well as the
rate of ethyl acetate formation for the rst hour of reaction. In
spite of the high activity shown in the 1st run, the catalytic perfor-
mance of CX-S1 markedly decreases when reused in consecutive
runs, more pronounced from the 1st to the 2nd runs. After the
3rd cycle, the material was characterized by TPD. The amount of
SO
2
released, corresponding to the desorption of sulphonic groups,
had decreased from1623mol g
1
to 1024mol g
1
, which could
explain the decrease in catalytic activity. On the other hand, the
decrease in catalytic performance can also be explained by forma-
tion of sulphonic esters, as reported by Fraile et al. [30]. According
to their work, the formation of sulphonate esters accounts for
the deactivation behaviour in reactions taking place in alcohol
solvents.
4. Conclusions
Oxidation of carbon nanotubes and xerogels with nitric acid
increases their surface acidity, mainly as a result of the carboxylic
acid groups formed. Sulphonic acid groups are incorporated by
treatment with sulphuric acid, the resulting materials showing a
strong acidic character. These carbon materials exhibited excellent
performances as solid acid catalysts in the esterication of acetic
acid. The sulphonic acid groups seemto be more adequate for this
reaction than the oxygen containing groups incorporated by oxi-
dation with nitric acid. A good correlation was obtained between
the rate of formation of ethyl acetate and the concentration of sul-
phonic acid groups in the case of sulphuric acid treated samples.
Acknowledgments
This work was supported by projects: FREECATS nanced by
the European Union 7th FP (2007-2013), Grant No. 280658; PEst-
C/EQB/LA0020/2011 and PTDC/EQU-ERQ/101456/2008, nanced
by FEDER through COMPETE Programa Operacional Factores de
Competitividade and by FCT Fundac o para a Cincia e a Tecnolo-
gia; and NORTE-01-0162-FEDER-000051 (SAIECT-IEC/2/2010).
References
[1] R.W. Coughlin, Carbon as adsorbent and catalyst, Product R&D 8 (1969) 1223.
[2] J.L. Figueiredo, M.F.R. Pereira, The role of surface chemistry in catalysis with
carbons, Catalysis Today 150 (2010) 27.
[3] P. Serp, J.L. Figueiredo, Carbon Materials for Catalysis, John Wiley &Sons, Hobo-
ken, NJ, 2009.
[4] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. rfo, Modication of the
surface chemistry of activated carbons, Carbon 37 (1999) 13791389.
[5] N. Mahata, M.F.R. Pereira, F. Surez-Garca, A. Martnez-Alonso, J.M.D. Tascn,
J.L. Figueiredo, Tuning of texture and surface chemistry of carbon xerogels,
Journal of Colloid and Interface Science 324 (2008) 150155.
[6] A.G. Gonc alves, J.L. Figueiredo, J.J.M. rfo, M.F.R. Pereira, Inuence of the sur-
face chemistry of multi-walled carbon nanotubes on their activity as ozonation
catalysts, Carbon 48 (2010) 43694381.
[7] Y.-C. Chiang, W.-H. Lin, Y.-C. Chang, The inuence of treatment duration
on multi-walled carbon nanotubes functionalized by H
2
SO
4
/HNO
3
oxidation,
Applied Surface Science 257 (2011) 24012410.
[8] V. Datsyuk, M. Kalyva, K. Papagelis, J. Parthenios, D. Tasis, A. Siokou, I. Kallitsis,
C. Galiotis, Chemical oxidation of multiwalled carbon nanotubes, Carbon 46
(2008) 833840.
[9] H.T. Gomes, S.M. Miranda, M.J. Sampaio, J.L. Figueiredo, A.M.T. Silva, J.L. Faria,
The role of activated carbons functionalized with thiol and sulfonic acid groups
in catalytic wet peroxide oxidation, Applied Catalysis B: Environmental 106
(2011) 390397.
[10] F. Peng, L. Zhang, H. Wang, P. Lv, H. Yu, Sulfonated carbon nanotubes as a strong
protonic acid catalyst, Carbon 43 (2005) 24052408.
[11] R. Liu, X. Wang, X. Zhao, P. Feng, Sulfonated ordered mesoporous carbon for
catalytic preparation of biodiesel, Carbon 46 (2008) 16641669.
[12] H. Yu, Y. Jin, Z. Li, F. Peng, H. Wang, Synthesis and characterization of sulfonated
single-walled carbon nanotubes and their performance as solid acid catalyst,
Journal of Solid State Chemistry 181 (2008) 432438.
[13] H.T. Gomes, S.M. Miranda, M.J. Sampaio, A.M.T. Silva, J.L. Faria, Activated car-
bons treatedwithsulphuric acid: catalysts for catalytic wet peroxide oxidation,
Catalysis Today 151 (2010) 153158.
[14] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. rfo, Characterization of
active sites on carbon catalysts, Industrial and Engineering Chemistry Research
46 (2007) 41104115.
[15] M. Hara, T. Yoshida, A. Takagaki, T. Takata, J.N. Kondo, S. Hayashi, K. Domen, A
carbon material as a strong protonic acid, Angewandte Chemie International
Edition 43 (2004) 29552958.
[16] M. Okamura, A. Takagaki, M. Toda, J.N. Kondo, K. Domen, T. Tatsumi, M. Hara,
S. Hayashi, Acid-catalyzed reactions on exible polycyclic aromatic carbon in
amorphous carbon, Chemistry of Materials 18 (2006) 30393045.
[17] V.L. Budarin, J.H. Clark, R. Luque, D.J. Macquarrie, Versatile mesoporous car-
bonaceous materials for acid catalysis, Chemical Communications (2007)
634636.
[18] N. Job, R. Pirard, J. Marien, J.-P. Pirard, Porous carbon xerogels with texture
tailored by pH control during solgel process, Carbon 42 (2004) 619628.
[19] R.R.N. Marques, B.F. Machado, J.L. Faria, A.M.T. Silva, Controlled generation of
oxygen functionalities on the surface of single-walled carbon nanotubes by
HNO
3
hydrothermal oxidation, Carbon 48 (2010) 15151523.
[20] I.D. Rosca, F. Watari, M. Uo, T. Akasaka, Oxidation of multiwalled carbon nano-
tubes by nitric acid, Carbon 43 (2005) 31243131.
56 R.P. Rocha et al. / Catalysis Today 218219 (2013) 5156
[21] R.P. Rocha, A.M.T. Silva, G. Dra zi c, M.F.R. Pereira, J.L. Figueiredo, Supported Pt-
particles on multi-walled carbon nanotubes with controlled surface chemistry,
Materials Letters 66 (2012) 6467.
[22] G. Ovejero, J.L. Sotelo, M.D. Romero, A. Rodrguez, M.A. Oca na, G. Rodrguez,
J. Garca, Multiwalled carbon nanotubes for liquid-phase oxidation. Function-
alization, characterization, and catalytic activity, Industrial and Engineering
Chemistry Research 45 (2006) 22062212.
[23] M. Monthioux, B.W. Smith, B. Burteaux, A. Claye, J.E. Fischer, D.E. Luzzi, Sen-
sitivity of single-wall carbon nanotubes to chemical processing: an electron
microscopy investigation, Carbon 39 (2001) 12511272.
[24] A.M.T. Silva, B.F. Machado, J.L. Figueiredo, J.L. Faria, Controlling the surface
chemistry of carbon xerogels using HNO
3
-hydrothermal oxidation, Carbon 47
(2009) 16701679.
[25] E. Cano-Serrano, G. Blanco-Brieva, J.M. Campos-Martin, J.L.G. Fierro, Acid-
functionalized amorphous silica by chemical grafting quantitative oxidation
of thiol groups, Langmuir 19 (2003) 76217627.
[26] J.G.C. Shen, R.G. Herman, K. Klier, Sulfonic acid-functionalized mesoporous sil-
ica: synthesis, characterization, and catalytic reaction of alcohol coupling to
ethers, Journal of Physical Chemistry B 106 (2002) 99759978.
[27] Q. Yang, J. Liu, J. Yang, M.P. Kapoor, S. Inagaki, C. Li, Synthesis, characteriza-
tion, and catalytic activity of sulfonic acid-functionalized periodic mesoporous
organosilicas, Journal of Catalysis 228 (2004) 265272.
[28] R.P. Rocha, J.P.S. Sousa, A.M.T. Silva, M.F.R. Pereira, J.L. Figueiredo, Catalytic
activity and stability of multiwalled carbon nanotubes in catalytic wet air
oxidation of oxalic acid: the role of the basic nature induced by the surface
chemistry, Applied Catalysis B: Environmental 104 (2011) 330336.
[29] A.P. Terzyk, Further insights into the role of carbon surface functionalities in
the mechanism of phenol adsorption, Journal of Colloid and Interface Science
268 (2003) 301329.
[30] J.M. Fraile, E. Garca-Bordej, L. Roldn, Deactivation of sulfonated hydro-
thermal carbons in the presence of alcohols: evidences for sulfonic esters
formation, Journal of Catalysis 289 (2012) 7379.

Das könnte Ihnen auch gefallen