Controlled Surface Functionalization of Multiwall Carbon Nanotubes by HNO3 Hydrothermal Oxidation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

Available at www.sciencedirect.com

ScienceDirect

journal homepage: www.elsevier.com/locate/carbon

Controlled surface functionalization of multiwall


carbon nanotubes by HNO3 hydrothermal oxidation

Vlassis Likodimos a,*, Theodore A. Steriotis a, Sergios K. Papageorgiou a,


George Em. Romanos a, Rita R.N. Marques b, Raquel P. Rocha b, Joaquim L. Faria b,
Manuel F.R. Pereira b, José L. Figueiredo b, Adrián M.T. Silva b,*, Polycarpos Falaras a,*
a
Institute of Advanced Materials, Physicochemical Processes, Nanotechnology and Microsystems, National Center for Scientific Research
‘‘Demokritos’’, 153 10 Aghia Paraskevi Attikis, Athens, Greece
b
LCM – Laboratory of Catalysis and Materials – Associate Laboratory LSRE/LCM, Faculdade de Engenharia, Universidade do Porto,
Rua Dr. Roberto Frias Q2 s/n, 4200-465 Porto, Portugal

A R T I C L E I N F O A B S T R A C T

Article history: Controlled surface functionalization is demonstrated by nitric acid hydrothermal oxidation
Received 29 August 2013 on multiwall carbon nanotubes (MWCNTs). The formation and evolution of oxygen func-
Accepted 10 December 2013 tional groups were systematically investigated as a function of the HNO3 concentration
Available online 16 December 2013 on MWCNTs with different structural and morphological characteristics, employing tem-
perature-programmed desorption coupled with mass spectrometry, thermogravimetry
and differential scanning calorimetry, Raman spectroscopy and N2 porosimetry analysis.
Hydrothermal treatment provides controlled MWCNT modification by specific oxygen
functionalities at amounts determined by the morphology, texture and crystallinity of
the pristine materials. Hydrothermal oxidation competes well with the harsh boiling nitric
acid treatment regarding the total amount of oxygen functionalities, while requiring much
lower amounts of oxidizing agent and, most importantly, reducing amorphous carbon
deposits on the MWCNT surface, a major drawback of aggressive liquid phase oxidation
methods. Detailed pore structure analysis revealed a progressive increase of the surface
area upon hydrothermal functionalization, whereas the mesopore structure varied consis-
tently with the intrinsic MWCNT properties related to the packing of the nanotube bundles
and the reduction of amorphous carbon. These advantageous features render nitric acid
hydrothermal oxidation an efficient functionalization process to fine tune and optimize
the surface chemistry of MWCNTs for target applications, circumventing the need for addi-
tional purification post-processing.
 2013 Elsevier Ltd. All rights reserved.

1. Introduction lithium-ion battery electrodes to sensors and field emission


displays [1,2]. Recent developments in the mass production
Carbon nanotubes (CNTs) hold promise as one of the key of multiwall carbon nanotubes (MWCNTs), which has gradu-
materials in the rapidly emerging nanotechnology driven ally evolved from the bench to the large-scale manufacturing,
applications encompassing a broad range of potential prod- rendered them the material of choice for large volume appli-
ucts from conductive, high-strength composites and cations, in contrast to the prototype single wall carbon

* Corresponding authors.
E-mail addresses: likodimo@chem.demokritos.gr (V. Likodimos), adrian@fe.up.pt (A.M.T. Silva), papi@chem.demokritos.gr,
Robert_Hurt@brown.edu (P. Falaras).
0008-6223/$ - see front matter  2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.carbon.2013.12.030
312 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

nanotubes (SWCNTs), whose scale-up has lagged behind be- whose density correlated analytically with the HNO3 concen-
cause of their significantly higher cost [3]. Despite the marked tration. Detailed investigations by different experimental
progress in the synthesis and purification of CNTs [4], con- techniques confirmed that HNO3 hydrothermal oxidation en-
trolled functionalization of their hydrophobic and chemically dows controlled amounts of oxygen functionalities on
inert surface remains a major challenge for practical deploy- SWCNTs, while it inhibits extensive formation of carboxyl-
ment, especially for MWCNTs that being on the verge of com- ated carbonaceous fragments and harsh oxidative damages
mercialization attract particular interest [5]. Liquid phase of the tube structure that would otherwise require additional,
oxidation is the most prevalent process for the purification chemical consuming, purification steps [23].
of as-grown CNTs by removing amorphous carbon and resid- In this work, the potential of nitric acid hydrothermal oxi-
ual metal catalyst impurities [4] as well as for the modifica- dation was systematically explored for the controlled surface
tion of their surface chemistry by the covalent attachment functionalization of MWCNTs using two types of commer-
of oxygen-containing groups i.e. carboxylic acids and anhy- cially available materials grown by CCVD with largely differ-
drides, phenols, carbonyl quinones, and lactones, preferen- ent morphological characteristics (diameter and length). A
tially at the tube ends and CNT side-walls [6–8]. These combination of experimental techniques including tempera-
hydrophilic surface groups may drastically promote the CNTs’ ture-programmed desorption coupled with mass spectrome-
dispersability in various solvents and their colloidal stability try (TPD–MS), thermogravimetry and differential scanning
[9–11], enhance the CNTs’ interfacial coupling with polymer calorimetry (TGA–DSC), porosimetry analysis and Raman
matrices for the development of advanced nanocomposites spectroscopy were employed to quantify the nature, amount
[12], while simultaneously serving as anchoring sites for fur- and evolution of oxygen functional groups on the MWCNTs
ther chemical functionalization and processing [13]. as a function of the HNO3 concentration. The process effi-
Nitric acid is the most common reagent for the oxidation of ciency was evaluated in comparison with the aggressive oxi-
CNTs, which can selectively remove amorphous carbon and dation of MWCNTs in concentrated HNO3 at boiling
metal catalysts and at the same time generate abundant oxy- temperature together with the dependence of the extent of
genated groups at the exposed CNT surfaces, following a surface functionalization on the structural properties of the
sequential oxidation path that sets out at the highly reactive pristine MWCNT materials.
defect sites of CNTs [14]. However, HNO3 treatment, most fre-
quently performed under boiling conditions, is accompanied 2. Experimental section
by severe degradation effects, especially for SWCNTs, includ-
ing material’s loss and selective removal of metallic nano- 2.1. Materials
tubes, tube shortening as well as the formation of structural
defects and carbonaceous debris [15–18]. Recently, the genera- Pristine MWCNTs of large diameter (25–40 nm) produced by
tion of exfoliated carboxylated carbonaceous fragments via CCVD were purchased from Nanothinx S.A. with purity of
the nitric acid over-oxidation of SWCNTs and subsequent base 98.35% and length >10 lm, hereafter designated with the
washing, arose considerable concerns as to the process effec- product code NTX3 (http://www.nanotubesx.com). Compari-
tiveness, which could undermine the nanotubes’ dispersabili- son was made with thin MWCNTs purchased from NANO-
ty in different host matrices [19–24]. Although MWCNTs were CYLTM, designated as NC3100. These MWCNTs are produced
found to be more resistant to oxidation than SWCNTs [20], via the CCVD process with average diameter and length of
similar detrimental effects including the formation of oxida- 9.5 nm and 1.5 lm, respectively, and purity greater than 95%
tion debris by nitric acid reflux were also identified [25–30], (http://www.nanocyl.com).
which could largely compromise their performance for e.g.
catalytic [31] or electrochemical applications [32]. 2.2. Surface functionalization
Even though significant insights on the diverse HNO3 oxi-
dation pathways have been recently provided by the combi- Hydrothermal functionalization of the pristine NTX3 and
nation of thorough experimental and theoretical NC3100 samples was performed in a temperature controlled
investigations [14], kinetic control of the MWCNTs surface 160 mL autoclave (Parr Instruments) using HNO3 aqueous
functionalization by the oxidation duration and/or reaction solutions with variable concentrations (0.05, 0.10, 0.20 and
temperature remains a challenging task [20], justifying the 0.30 mol L1) at 473 K. For NTX3, 0.2 g of MWCNTs were added
high interest in the development of controllable purifica- to 75 mL of a HNO3 aqueous solution with the desired concen-
tion/functionalization processes of MWCNTs [33–35]. This is tration, the vessel was sealed and the solution was flushed
particularly important for commercial MWCNTs produced with nitrogen for 5 min to remove dissolved oxygen. Then,
by catalytic chemical vapor deposition (CCVD) [36,37], the the system was pressurized with 0.5 MPa of nitrogen and
most established technique for large scale production, whose heated up to 473 K at autogeneous pressure under continuous
structural properties may be significantly impaired by the stirring at 300 rpm. After 2 h of operation, the MWCNTs were
soot-like amorphous carbon emerging as pyrolysis by-product recovered, washed several times with distilled water until a
[34] and inherent defects in the graphitic tube shells due to neutral pH of the rinsing water and dried overnight at 393 K.
the relatively low growth temperature [38]. Recently, nitric A blank experiment was also performed using distilled water
acid hydrothermal oxidation was developed as a controllable instead of nitric acid. The same procedure was used to func-
and mild functionalization method to tailor the surface tionalize NC3100 samples, but in that case a higher loading
chemistry of carbon xerogels [39] and SWCNTs [8] by the (0.5 g) of MWCNTs was employed in order to allow compari-
thermally activated generation of oxygen functional groups, son at different materials’ loads. For comparison, nitric acid
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 313

oxidation was also performed by direct contact of the centrations as for the NTX3, though using higher amount
MWCNTs with HNO3 at boiling temperature. In a typical acti- (0.5 g) of NC3100 material (Fig. S1 in the Supplementary Infor-
vation run, 0.5 g or 2.0 g of the pristine NTX3 or NC3100 mate- mation). To quantify the relationship between oxygenated
rials, respectively, were immersed in 150 mL of a HNO3 groups and HNO3 concentrations, the total amount of surface
solution with a concentration of 7.0 M. A round bottom flask groups released as CO and CO2 was determined from the area
equipped with a condenser was used and the suspension under the corresponding TPD spectra (Table 1).
was heated to boiling temperature and kept under magnetic The weight percentage of the total amount of molecular
stirring for 3 h. After cooling, the suspension was washed oxygen released as CO and CO2 (wt.% O2), which is represen-
up to a neutral pH of the rinsing water was attained and the tative of the total amount of oxygenated groups introduced
recovered nanotubes were dried overnight at 393 K. on the CNT surface, was calculated based on the amounts
determined by the TPD analysis. Both pristine NTX3 and
2.3. Materials characterization NC3100 materials comprised low amounts of surface groups
leading to similar CO/CO2 ratios and a low amount of released
The surface chemistry of the MWCNTs was quantified by tem- wt.% O2. Furthermore, hydrothermal treatment without nitric
perature programmed desorption (TPD) analysis using an AMI- acid was verified to cause negligible effects on the poor sur-
200 Catalyst Characterization Instrument (Altamira Instru- face chemistry of the pristine NTX3 MWCNTs. On the other
ments) equipped with a quadruple mass spectrometer (Ame- hand, hydrothermal treatment in the presence of nitric acid
tek, Mod. Dymaxion). The sample (0.1 g) was placed in a U- resulted in a marked increase of the release of both CO and
shaped quartz tube and heated at 5 K min1 in an electrical fur- CO2 that depended strongly on the HNO3 concentration for
nace under a constant flow of 25 cm3 min1 of helium, used as both NTX3 and NC3100 MWCNTs, as shown in Fig. 2. Previous
carrier gas. The amount of CO and CO2 released was deter- studies [8] have pointed out that oxygen containing groups
mined using the calibration performed at the end of each anal- are not created on the surface of SWCNTs when hydrother-
ysis and the experimental error was lower than 8%. mal oxidation is performed at a lower temperature (393 K),
The materials texture was determined by N2 adsorption– the two key parameters of the functionalization process being
desorption isotherms at 77 K in the relative pressure range the temperature and HNO3 concentration. However, pressure
105–0.995 on a Quantachrome Autosorb-1-MP apparatus. during hydrothermal treatment may also promote surface
Each sample (60–80 mg) was outgassed under high vacuum group kinetics and the homogenous generation of oxygen
at 573 K overnight, prior to analysis. Thermal analysis studies functional groups within the MWCNTs layers, as inferred
were conducted on a Setaram SETSYS Evolution 16/18, TGA/ from their low thermal stability (Section 3.2). In addition, a
DSC analyzer. For each experiment, approximately 20 mg of single exponential function was found to describe analyti-
the material were loaded on the sample crucible, the system cally the evolution of CO and CO2 amounts with the HNO3
was sealed, fed with 16 cm3/min air (purity 99.9%) and a tem- concentration, in qualitative agreement with the results ob-
perature ramp (10 K/min) was applied. For simultaneous DSC tained for hydrothermally treated SWCNTs [8,23], and carbon
measurements the system was pre-calibrated with a series xerogels [39]. This shows that oxygen surface functionaliza-
of standards following the manufacturer’s procedure. tion of both MWCNTs and SWCNTs can be accurately con-
Raman measurements were performed in backscattering trolled through the hydrothermal methodology regardless of
configuration using a Renishaw inVia Reflex microscope with the structural differences between them and the different
an Ar+ ion laser (k = 514.5 nm) and a high power near infrared amounts of surface groups introduced on the CNT’s surface.
(NIR) diode laser (k = 785 nm) as excitation sources. The laser However, despite the qualitatively similar evolution of
light was focused on the samples using a 50· objective lens of oxygen functionalities on the hydrothermally treated materi-
a Leica DMLM microscope at power density lower than als, the degree of oxygen functionalization as monitored by
0.05 mW/lm2 for both laser lines, to avoid sample heating. the CO, CO2 and CO/CO2 dependence on [HNO3], varied appre-
Spectral deconvolution was carried out by non-linear least ciably for the two different types of MWCNTs. In fact, the
square fitting of the Raman peaks to a mixture of Lorentzian amounts of oxygenated groups released as CO and CO2 for
and Gaussian lineshapes. NC3100 exceeded significantly the corresponding ones of
NTX3 at HNO3 concentrations above 0.1 mol L1 (Table 1). This
difference is further augmented if we take into account the
3. Results and discussion higher loading of NC3100 MWCNTs (0.5 g) that would be ex-
pected to effectively decrease the amount of oxygen function-
3.1. Identification and quantification of oxygenated alities on the MWCNT surface at the same HNO3
groups concentration. These distinct differences between NTX3 and
NC3100, indicate that the nanotube morphology and particu-
Fig. 1 displays the TPD spectra arising from specific oxygen- larly the MWCNT’s diameter and length, which differ consid-
ated groups released as CO and CO2 from the surface of the erably between the two types of CCVD tubes, are critical
hydrothermally treated NTX3 MWCNTs as a function of the parameters for the extent of surface functionalization by
HNO3 concentration. The TPD profiles measured for the pris- the HNO3 hydrothermal process.
tine NTX3 material and for the blank experiment in the ab- Furthermore, comparison with the corresponding results
sence of HNO3 are also included for comparison. on hydrothermally treated SWCNTs [8] shows that the degree
Qualitatively similar TDP spectra were acquired for the of surface functionalization of the MWCNTs is systematically
thin and narrow NC3100 MWCNTs under identical HNO3 con- lower, especially for NTX3, where the amount of groups re-
314 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

0.6
(a)

CO (µ mol g s )
-1
0.4

-1
0.2

0.0

-1
0.3

)
L
0.2

ol
m
0.1

](
0.05

O
3
0 (Blank)

N
[H
0 - Pristine
300 400 500 600 700 800 900 1000 1100 1200 1300

T (K)
0.3

(b)

CO2 (µ mol g s )
-1
0.2

-1
0.1

0.0

-1
0.3

)
L
0.2

ol
0.1 (m
3]
0.05
O

0 (Blank)
N
[H

0 - Pristine
300 400 500 600 700 800 900 1000 1100 1200 1300

T (K)

Fig. 1 – TPD spectra for the pristine and hydrothermally treated NTX3 MWCNTs at different HNO3 concentrations and 473 K:
(a) CO and (b) CO2 release.

Table 1 – Total amount of CO, CO2 and wt.% O2 calculated from the TPD spectra obtained for the hydrothermally functionalized
NTX3 and NC3100 at different HNO3 concentrations and 473 K.
[HNO3](mol L1) NTX3 NC3100
CO (±20) CO2 (±20) CO/CO2 O2 (wt.%) CO (±20) (lmol g1) CO2 (±20) (lmol g1) CO/CO2 O2 (wt.%)
(lmol g ) (lmol g1)
1

0 – pristine 218 43 5.12 0.5 178 33 5.32 0.4


0 (Blank) 278 23 12.3 0.5 – – – –
0.05 946 304 3.11 2.5 489 195 2.51 1.4
0.10 1156 482 2.40 3.4 973 388 2.51 2.8
0.20 1463 545 2.68 4.1 1742 588 2.96 4.7
0.30 1573 662 2.38 4.6 2015 680 2.96 5.4
7.0 (Boiling) 1344 751 1.79 4.6 1511 767 1.97 4.9

leased as CO is nearly two times lower for all the HNO3 con- ing that the generation of surface groups released as CO2
centrations. Although the TPD profiles of the hydrothermally with respect to CO is favored on the MWCNTs surface in com-
treated MWCNTs are similar to the corresponding ones of parison with SWCNTs. In typical liquid phase functionaliza-
SWCNTs, the CO/CO2 ratios for the MWCNTs (<3.11) are con- tion methods with acids, the CO/CO2 ratio tends to decrease
siderably lower than those of the SWCNTs (>3.50) [8], imply- with the increase of the concentration of the oxidizing agent
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 315

2500 1500 dence of the amount of molecular O2 was thus found to vary
NTX3
CO
NTX3
= 1583 - 1290 exp (- 12.4 [HNO3])
considerably for the different types of CNTs, the most pro-
2000 NC3100 nounced differences occurring between NC3100 and NTX3.
NC3100
CO = 3186 - 3071 exp (- 3.4 [HNO3]) In particular, the short and thin (length of 1.5 lm and diam-
1000
eter of 10 nm) NC3100 MWCNTs were functionalized to a
CO (µ mol g )

CO2 (µ mol g )
-1

-1
1500
considerably higher extent than the much longer (>10 lm)
and thicker (diameter in the range of 25–40 nm) NTX3 nano-
1000
500
tubes, in accordance with the respective BET surface area of
NTX3 the pristine materials (SBET  300 and 80 m2 g1 for NC3100
(NTX3)
500 CO2 =650-625 exp(-11.9 [HNO3])
and NTX3, respectively), i.e. a higher functionalization is ob-
NC3100
CO2
NC3100
=837-815 exp(-5.7 [HNO3])
tained for MWCNTs with higher surface area. This confirms
0 0 that the CNTs morphology and especially the MWCNT’s
0.00 0.05 0.10 0.15 0.20 0.25 0.30
diameter and length and the concomitant variation of the
-1
[HNO3] (mol L ) surface area, dictate to a large extent the degree of surface
functionalization. Additionally, regarding both NTX3 and
Fig. 2 – Evolution of the amount of CO and CO2 released from
SWCNTs, Fig. 3 shows a plateau for the wt.% O2 evolution at
the surface of the hydrothermally treated NTX3 and NC3100
the higher [HNO3]/mCNT ratios, which was not observed for
MWCNTs vs. the HNO3 concentration. (A color version of
the NC3100 samples since only lower [HNO3]/mCNT ratios were
this figure can be viewed online.)
tested in this particular case.
The total amount of CO and CO2, as well as the [HNO3]/
mCNT ratio, were also normalized to the BET surface area of
due to the introduction of more acidic groups on the CNT sur- each sample (i.e. CO + CO2/SBET vs. [HNO3]/mCNT/SBET, not
face that are released as CO2. This implies the formation of shown). The amount of groups per surface area were thus
relatively more acidic groups, in comparison to basic groups, found to be slightly higher for SWCNTs in comparison to
on the surface of MWCNTs than on SWCNTs oxidized under NC3100, corroborating that the surface of the longer and thin-
similar conditions. ner SWCNTs is more easily oxidized than that of the shorter
In order to further assess the interplay between the effi- and thicker MWCNT NC3100. However, the CO + CO2/SBET ra-
ciency of hydrothermal oxidation for CNT surface functional- tio for a given [HNO3]/mCNT/SBET was higher for NTX3 than
ization and the intrinsic materials properties, the total for the other two materials, even if the total amount of groups
amount of molecular oxygen released from the materials’ sur- introduced in these MWCNTs is much lower (Fig. 3), indicat-
face (wt.% O2) was compared for both types of MWCNTs, ing that NTX3 offers less (but more reactive) sites for surface
NTX3 and NC3100, including previous results on hydrother- functionalization.
mally treated SWCNTs (SBET  400 m2 g1 for SWCNTs) [8] as To explore further the modification of the MWCNT’s sur-
a function of the [HNO3]/mCNT ratio, which is independent face chemistry by hydrothermal oxidation, the nature of the
of the CNTs loading (Fig. 3). In fact, it has been proved that distinct functional groups and their amounts were deter-
the wt.% O2 for two different loadings of SWCNTs (namely mined as function of the HNO3 concentration by deconvolu-
0.2 and 0.5 g) scale consistently when plotted as a function tion of the CO and CO2 TPD spectra following the procedure
of the [HNO3]/mCNT ratio [8], which may thus serve as com- previously developed for activated carbons [40,41]. All spectra
mon scale for the comparison of functionalized CNTs by the were deconvoluted and the amount of each group (deter-
hydrothermal oxidation methodology. The [HNO3] depen- mined by calculating the area under the correspondent peak)
is presented as a function of the HNO3 concentration in Fig. 4.
Phenols were the main groups released as CO and carboxylic
acids those released as CO2, for both types of MWCNTs. Car-
10
SWNTs bonyls/quinones and lactones were also identified by the
O2 = 8.46 - 6.74 exp (- 2.04 [HNO3])
MWNTs-NC3100 deconvolution of the CO and CO2 spectra, respectively, while
MWNTs-NTX3 r2 = 0.985
8 carboxylic anhydrides were identified on both CO and CO2
spectra. In addition, an exponential analytic function was
O2 (wt. %)

6 O2 = 7.55 - 7.29 exp (- 2.14 [HNO3]) found to fit well the [HNO3] dependence of the amount of
r2 = 0.981
each single oxygenated moiety (not shown), as found for
4 the total amount of groups on the surface of MWCNTs
(Fig. 2). The amount of all identified surface groups increased
O2 = 4.60 - 4.09 exp (- 2.46 [HNO3])
2
with the HNO3 concentration except for carbonyl/quinone
r2 = 0.989
functionalities. The proposed mechanism for the creation/
evolution of oxygenated surface groups on CNTs is based on
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 a progressive pathway where the carbonyl/quinone groups
[HNO3]/mCNT (mol L g ) -1 -1 (C@O) are formed and further transformed into phenolic
(AOH) and carboxylic (ACOOH) functionalities [14,16,42,43].
Fig. 3 – Amount of molecular oxygen present as oxygenated Under the conditions applied in the present work, the amount
groups in the surface of SWCNT, NC3100 and NTX3 as a of carbonyl/quinones functionalities decreased with the
function of the [HNO3]/mCNT ratio. increase of the HNO3 concentration for both NTX3 and
316 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

(a) 300
Carboxylic Anhydrides
1200 (b) 500
Carboxylic Acids
Phenols Carboxylic Anhydrides
250 Carbonyl / Quinones 400 Lactones
900

CO2 (µ mol g )
-1
CO (µ mol g )
CO (µ mol g ) 200

-1
-1

300
150 600
200
100
300
50
100

0 0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30
-1 -1
[HNO3] (mol L ) [HNO3] (mol L )

(c) 500 Carboxylic Anhydrides


2000 (d) 600
Carboxylic Acids
Phenols 500 Carboxylic Anhydrides
400 Carbonyl / Quinones Lactones
1500

CO2 (µmol g-1)


CO (µmol g-1)
CO (µmol g-1)

400
300
1000 300
200
200
500
100
100

0 0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30
[HNO3] (mol L ) -1
[HNO3] (mol L-1)

Fig. 4 – Evolution of the amount of specific oxygenated groups created at the surface of the hydrothermally treated NTX3 (a
and b) and NC3100 (c and d) MWCNTs with the HNO3 concentration and released as CO (a and c) and CO2 (b and d).

NC3100 MWCNTs, suggesting that the development of pheno- with respect to SWCNTs. This could effectively alter the nat-
lic and carboxylic surface groups could in fact result from the ure and density of reactive sites for oxygen group formation
conversion of carbonyl/quinones initially created on the and lead to diverse oxidations mechanism under hydrother-
MWCNTs’ surface. mal conditions. Recent theoretical work on the oxidation of
Furthermore, comparison between the two types of MWCNTs under HNO3 reflux predicted a four step reaction
MWCNT showed that the generation of both phenols and car- mechanism for a monovacancy defect leading to a surface
boxylic groups was markedly enhanced for NC3100, confirm- carboxylic group accompanied with vacancy enlargement
ing that this material offers more sites for surface [14]. On the other hand, a different oxidation mechanism
functionalization compared with the long and large diameter may occur for SWNTs, where the confined diameter and di-
NTX3 MWCNTs. Moreover, the amount of carboxylic anhy- verse chirality were predicted to lower the adsorption barrier
drides for MWCNTs was comparable (NTX3) or even higher for NO2 attack on the tube walls leading to the preferential
(NC3100) than that estimated for SWCNTs at the same condi- functionalization and even selective removal of small diame-
tions [8]. Regarding the other surface groups, the amounts of ter metallic SWCNTs [17].
carbonyl/quinones, phenols and lactones were smaller in For comparison, the boiling acid method was also applied
MWCNTs compared to the corresponding ones obtained for for the surface functionalization of the NTX3 and NC3100
SWCNTs. However, the amount of carboxylic acids was simi- MWCNTs. Fig. 5 shows the TPD spectra for NTX3 treated with
lar for NTX3 or considerably larger for the NC3100 MWCNTs 7.0 mol L1 HNO3 at boiling temperature and their deconvolu-
in comparison with the SWCNTs, verifying that the transfor- tion to the individual components of the different oxygenated
mation of carbonyl/quinones to carboxylic acids is more surface groups (see Fig. S2 in the Supplementary Information
favorable on the surface of MWCNTs than on SWCNTs, in par- for the corresponding spectra of NC3100 MWCNTs).
ticular when compared to phenols. This justifies the lower The corresponding amount of surface groups released as
CO/CO2 ratios found for hydrothermally treated MWCNTs CO and CO2 as well as wt.% O2 are included in Table 1. The to-
with respect to their SWCNT analogues and further empha- tal amount of surface groups was comparable, though sys-
sizes the high efficiency of hydrothermal oxidation for CCVD tematically lower, than that obtained under hydrothermal
MWCNTs under optimal conditions. The enhanced surface treatment at the highest HNO3 concentration (0.3 mol L1),
reactivity of hydrothermally treated MWCNTs for ACOOH especially for the highly functionalized NC3100 samples.
generation may be in principle rationalized by the inherent These results indicate that apart from being a controllable
differences in the spatial dimensions and defect structure functionalization process that consumes a lower amount of
(e.g. CNT caps with pentagons and heptagons) of MWCNTs oxidizing agent, hydrothermal oxidation competes well with
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 317

temperatures, where the carboxylic acids evolve in the CO2


(a) 0.4
PH spectrum (identified as CAc in Fig. 5a), whose origin, however,
is not fully understood at present [41].
0.3
An additional process that occurs during the hydrothermal
oxidation is the weight loss (WL) that corresponds to the dif-
CO (µ mol g-1s-1)

0.2 CAn ference between the weight of CNTs introduced into the auto-
CQ clave and the final weight of the recovered material after the
CAc hydrothermal run. Fig. 6 compares the observed WL (%) as a
0.1
function of the [HNO3]/mCNT ratio for the SWCNTs, NTX3
and NC3100 MWCNTs. The observed WL variation can be fit-
0.0 ted to exponential functions, resembling the dependence of
300 400 500 600 700 800 900 1000 1100 1200 1300
the oxygenated groups with the [HNO3]/mCNT ratio.
T (K)
This effect can be associated with the preferential forma-
tion of oxygenated groups on carboxylated carbonaceous
(b) 0.25
fragments (CCFs) produced by the CNT’s consumption during
SA HNO3 treatment, especially SWCNTs that are most suscepti-
0.20
WA ble to acid oxidation [20–24]. It was proposed that under
strong enough oxidizing conditions the adsorbed CCFs on
CO2 (µ mol g-1s-1)

0.15 CAn
the CNT walls can be fully gasified to CO2, accounting for
0.10 the observed WL and its dependence on the HNO3 concentra-
LC
tion [8]. In the present case, the WL increased upon hydro-
0.05
thermal treatment at 0.30 mol L1 HNO3 up to around 50%,
25% and 30% for SWCNTs, NTX3 and NC3100 materials,
0.00
respectively. These results confirm the higher stability of the
300 400 500 600 700 800 900 1000 1100 1200 1300 MWCNT structure to nitric acid oxidation compared with
T (K)
the SWCNTs. Moreover, NC3100 exhibited appreciable WL
Fig. 5 – Deconvolution of TPD spectra for NTX3 treated with upon hydrothermal oxidation, approaching that of SWCNTs,
[HNO3] = 7.0 mol L1 at boiling temperature: groups released consistently with their highly functionalized state.
as (a) CO and (b) CO2 (PH – phenols; CAn – carboxylic
anhydrides; CQ – carbonyl quinones; LC – lactones; CAc – 3.2. Thermal analysis
carboxylic acids; SA – strong acidic CAc; WA – weakly acidic
CAc). Further evidence for the efficiency of MWCNT hydrothermal
oxidation was provided by TGA–DSC measurements. Fig. 7
compares the TGA curves and the corresponding enthalpy
the harsh boiling HNO3 treatment in terms of the generation (DSC) changes on the pristine and hydrothermally treated
of sufficient quantities of oxygen functionalities on the NTX3 and NC3100 MWCNTs at different HNO3 concentra-
MWCNT surface. Nevertheless, the CO2 amount was relatively tions. The effect of hydrothermal treatment was evident on
higher (by 13%) in the MWCNTs oxidized under boiling acid both classes of materials, as increasing the HNO3 concentra-
compared to the hydrothermally treated ones (Table 1). Spe- tion resulted in significantly lower thermal stability of the
cifically, the amounts of carboxylic acids for the NTX3 sample functionalized MWCNTs due to the existence of covalently
treated with the boiling method (7.0 mol L1) and for the
hydrothermally treated sample at the highest HNO3 concen-
tration (0.3 mol L1) reached 520 and 370 lmol g1, respec-
60
tively. In the case of NC3100, the corresponding amounts SWNTs
WL = 52.3 - 47.3 exp (- 1.52 [HNO3])
were 331 and 222 lmol g1, respectively, indicating that more 50
MWNTs-NC3100
r2 = 0.995
MWNTs-NTX3
acidic surface groups were produced on the MWCNT surface
not only due to the aggressive oxidizing conditions of the boil- 40
WL (%)

ing acid method but also because hydrothermal oxidation is


WL = 153 - 149 exp (- 0.30 [HNO3])
performed at 473 K, a temperature where a fraction of carbox- 30 r2 = 0.999
ylic acids can be partially released. In this context, the CO2
TPD spectrum of the hydrothermally treated samples 20
(Fig. 1b) differed appreciably from that obtained under boiling WL = 32.4 - 27.4 exp (- 1.00 [HNO3])
10 r2 = 0.999
conditions (Fig. 5b). In the latter case, CO2 evolved from lower
temperatures, due to the higher amount of carboxylic acids in
0
this sample, allowing the differentiation of two different 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
types of carboxylic acids by the deconvolution of the CO2 [HNO3]/mCNT (mol L -1 g-1)
TPD spectrum, namely strong acidic (SA) and weakly acidic
(WA) carboxylic acids, which have been established to evolve Fig. 6 – Weight loss (WL) observed after hydrothermal
at lower and higher temperatures, respectively [41]. A treatment of NTX3, NC3100 and SWCNTs as a function of
shoulder in the CO spectrum can be also traced at the same the [HNO3]/mCNT ratio.
318 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

150
100
120 NC3100

Heta flow (mW)


80 Pristine
Weight (%)
90 0.1 M
60 NC3100
0.3 M
Pristine 60
40 0.1 M
0.3 M 30
20
0
0
150
100
120 NTX3
Pristine

Heta flow (mW)


80
Weight (%)

90 0.1 M
60 NTX3 0.3 M
Pristine 60
40 0.1 M
0.3 M 30
20
0
0
300 400 500 600 700 800 900 1000 300 400 500 600 700 800 900 1000
Temperature (K) Temperature (K)
Fig. 7 – TGA–DSC curves on the pristine and hydrothermally treated NC3100 and NTX3MWCNTs at different HNO3
concentrations. (A color version of this figure can be viewed online.)

bonded oxidizing surface groups and the concurrent partial to thermal decomposition [5] most pertinent to the large
loss of crystallinity [44]. diameter NTX3 MWCNTs comprising a high number (25–50)
Specifically, for the pristine NC3100, thermal decomposi- of rolled graphitic layers. Furthermore, TGA measurements
tion initiated at 723 K completing at 968 K. On the other hand, on the NTX3 and NC3100 MWCNTs oxidized with 7.0 mol L1
for the functionalized samples thermal degradation started at HNO3 at boiling temperature (Fig. S3, Supplementary Infor-
603 and 543 K for the treated samples at 0.1 and 0.3 mol L1 mation), revealed considerably higher thermal stability of
HNO3, followed by complete burn out at 878 and 813 K, both samples compared to the corresponding (0.3 mol L1)
respectively, indicative of the drastic reduction of the hydrothermally functionalized ones, in spite of the presence
MWCNTs’ thermal stability by hydrothermal oxidation. Fur- of comparable amounts of oxygen functional groups (Table 1).
thermore, no marked differences were observed on the ther- This would further imply that hydrothermal treatment re-
mal stability between the pristine NTX3 and the NC3100 sults in more homogenous distribution of oxygenated groups
samples. The onset of thermal degradation for the pristine within the MWCNTs, most likely through the beneficial effect
NTX3 occurred at ca. 753 K, only 30 K higher than the corre- of pressure, leading to the distinct reduction of MWCNT ther-
sponding one of NC3100 MWCNTs, despite their much larger mal stability compared to the boiling acid method. In the lat-
diameter (the oxidation temperatures determined from the ter case, surface modification including the formation of
maxima of the corresponding mass loss rates were 861 and amorphous carbon (vide infra) may be limited to the external
870 K for the NTX3 and NC3100, respectively). This further im- layers of the MWCNTs, leaving intact the internal CNT struc-
plies that the thin and short NC3100 MWCNTs present a high ture and justifying the relatively higher thermal stability of
degree of crystallinity, in spite of their small diameter that boiling acid treated samples.
would otherwise lead to considerably lower oxidation temper-
ature and thermal stability [45]. 3.3. Raman spectroscopy
Moreover, hydrothermal functionalization reduced drasti-
cally the thermal stability of NTX3 with increasing HNO3 con- The effect of hydrothermal oxidation on the MWCNTs’ micro-
centration. However, contrary to the NC3100 MWCNTs, two structure was further investigated by Raman spectroscopy.
major components with different thermal properties were Fig. 8 presents the evolution of the Raman spectra for the
clearly differentiated by a slope change of the TGA traces hydrothermally functionalized NTX3 samples compared with
around 763 K and a double peak profile of the DSC curve for the pristine one at two different excitation wavelengths in the
the functionalized NTX3 samples. The first component was NIR (785 nm) and visible range (514.5 nm). The characteristic
identified at temperatures slightly below the corresponding Raman modes of MWCNTs were thus identified, the most
oxidation peak of NC3100, while the second one occurred at prominent being the first order tangential G band
higher temperatures, indicative of a MWCNT fraction more (1585 cm1) together with two highly dispersive modes,
resistant to oxidation. This distinctive behavior may be the defect-induced D band (1350 cm1 at 514.5 nm) and its
related to the diverse reactivity of inner and outer CNT layers overtone (G 0 band) at 2D (2700 cm1 at 514.5 nm). The D
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 319

successfully applied in several carbonaceous materials [53,54]


(a) 785 nm
D
Pristine including MWCNTs [34,55].
At the high frequency range (2300–3400 cm1), the Raman
Normalized Intensity
Pristine G
0.05 M D'
0.10 M
spectra could be well described by the superposition of the
D4
0.20 M weak non-dispersive step-like mode at 2450 cm1 related
0.30 M D3
to the overtone of the longitudinal optical phonon near the
K points of the two-dimensional graphite Brillouin zone with
1000 1200 1400 1600 1800 q = 0 [56], the G 0 and the defect-induced D + D 0 bands together
with a composite band at higher frequencies (Fig. S4, Supple-
mentary information). The latter band could be approximated
by two weak peaks at 3180 and 3240 cm1 corresponding to
800 1200 1600 2400 2800 3200 the frequencies of the G band overtone (2G) and its shifted
variant in graphitic materials due to the overbending of the
(b) graphite phonon dispersion curve [23,57].
514.5 nm G'
Fig. 9a and b show the [HNO3] dependence of the Raman
Normalized Intensity

intensity ratios of the D, D 0, D + D 0, D3 and D4 bands (ID/IG,


ID 0 /IG, ID+D 0 /IG, ID3/IG and ID4/IG) relative to the tangential G
mode for the NTX3 MWCNTs, calculated from the areas of
the corresponding Raman bands, at 514.5 and 785 nm, respec-
tively. A gradual decrease of the integrated intensity ratios for
D+D'
both the defect activated D, D 0, and D + D 0 modes and the two
2G
broad D3 and D4 bands due to amorphous/disordered carbon,
was evidenced at both excitation wavelengths upon succes-
800 1200 1600 2400 2800 3200 sive introduction of oxygen functionalities on the NTX3
Raman shift (cm )
-1 MWCNTs with increasing HNO3 concentration. This variation
apparently contrasts with the linear increment of the corre-
Fig. 8 – Raman spectra on the pristine and hydrothermally sponding defect-activated Raman band intensity ratios with
treated NTX3 MWCNTs at (a) 785 nm and (b) 514.5 nm. The the amount of oxygenated groups on hydrothermally func-
Raman intensity was normalized to that of the G band. The tionalized SWCNTs under the same conditions [23] as well
inset in (a) depicts representative results of the five-peak as the frequently observed increase of the ID/IG ratio with
spectral fitting for the pristine MWCNTs at 785 nm. (A color the reaction time for MWCNTs oxidized by the boiling acid
version of this figure can be viewed online.) method [25,58]. The decrease of the intensity ratios was fur-
ther accompanied by the progressive narrowing of the width
(full width at half-maximum-FWHM) of the defect activated
band originates from a double resonant Raman process that Raman bands, as shown in Fig. 9c, together with a systematic
involves elastic scattering from defects together with one- upward shift of their frequency reaching 5, 8 and 15 cm1 for
phonon inelastic scattering, rendering it the characteristic the D, D 0 and D + D 0 bands, respectively, under the highest
feature for the identification of structural disorder and de- HNO3 concentration of 0.30 mol L1.
fects in graphitic materials [46]. On the other hand, the sec- Both effects, i.e. the relative intensity reduction and the
ond-order G 0 band is activated independently of defects narrowing/shift of the defect activated Raman bands, point
through a two-phonon double resonance process [46], which to the preferential oxidation of the most reactive external
is responsible for its dispersive nature and sensitivity to the layer carbon layers on the surface of NTX3 MWCNTs with
structural ordering of the tube walls and the graphitic elec- hydrothermal treatment, in close analogy to previous results
tronic structure [47]. Two additional defect-induced modes on soot oxidation [53]. On the other hand, the G 0 band, which
of lower intensity were also observed: the D 0 band appearing is independent of defects and sensitive to the graphitic struc-
as a shoulder to the G mode at 1622 cm1 arising from a tural ordering [47], was significantly enhanced upon increas-
double resonance process similar to the D-band [46] and the ing the HNO3 concentration. Specifically, the intensity ratio
relatively broad D + D 0 combination mode (2950 cm1 at IG 0 /IG increased continuously with [HNO3], while the FWHM
514.5 nm) [48]. Most of the MWCNTs’ Raman bands exhibited of the G 0 band decreased drastically in the same concentra-
significant variations for the hydrothermally treated NTX3 tion range [Fig. 9d]. This effect can be rationalized by the par-
upon increasing the HNO3 concentration, at both excitation tial removal of the amorphous carbon deposits on the
energies. However, quantitative spectral analysis at the inter- MWCNTs or even the consumption of the most disordered/
mediate frequency range (1200–1800 cm1) in terms of the defective outer MWCNT shells upon hydrothermal oxidation
three independent Raman bands (G, D and D 0 ) resulted in [59], directly probed by the decrease of the D3 and D4 band
rather poor fitting results. Accurate spectral deconvolution intensities. The inner, more graphitic shells of the nanotubes
was obtained by including two additional broad bands, the are thus unveiled, which can be identified by the upsurge of
D3 (1500 cm1) and D4 (1200 cm1) ones (inset of Fig. 8a), re- the narrow G 0 band together with the relatively sharper and
lated to the presence of amorphous carbon and disordered upward shifted D, D 0, D + D 0 bands, whose intensity is primar-
graphitic material [49–51], respectively, following the ily determined by the inherent defect structure of the CCVD
five-peak fitting procedure of Sadezky et al. [52] that has been grown MWCNTs rather than the oxygen groups produced by
320 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

(a) (b) 4.5


2.0 514.5 nm 785 nm
4.0
1.6 ID/IG ID/IG
Intensity ratio

Intensity ratio
1.2 3.5

0.8 ID+D'/IG
1.5
ID4/IG
ID4/IG
0.4 1.0
ID'/IG ID'/IG
ID3/IG 0.5 ID3/IG
0.0

(c) 120 300 (d) 2.0 110


514.5 nm 514. 5 nm
D band 785 nm 250 1.6 785 nm
100

G' FWHM (cm )


-1
80
FWHM (cm )
-1

200 1.2
D+D' band
90

IG'/IG
40 150 0.4
D' band
80
100 0.3
30
50 0.2 70
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30
-1 -1
[HNO3] (mol L ) [HNO3] (mol L )

Fig. 9 – The HNO3 concentration dependence of the ID/IG, ID 0 /IG, ID+D 0 /IG, ID3/IG and ID4/IG integrated intensity ratios for the
hydrothermally functionalized NTX3 MWCNTs at (a) 514.5 and (b) 785 nm. Variation of the FWHM of the defect activated (D,
D 0 , D + D 0 ) Raman bands (c) and the intensity ratio IG 0 /IG and FWHM of the G 0 band (d) as a function of [HNO3]. (A color version
of this figure can be viewed online.)

(a) D
G D' 514.5 nm
Normalized Intensity

G'
D+D'
0.30 M 2G

0.20 M
0.10 M
0.05 M
Pristine

1200 1600 2400 2800 3200


-1
Raman shift (cm )

(b) (c)
ID/IG
2.0 140 D+D' band
FWHM (cm )

IG'/IG 120
-1
Intensity ratio

1.6 G' band


ID+D'/IG 100
0.8 D band
ID3/IG 60
ID'/IG
0.4 ID4/IG
40 D' band

20
0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3
-1 -1
[HNO3] (mol L ) [HNO3] (mol L )

Fig. 10 – (a) Raman spectra of the pristine and hydrothermally functionalized NC3100 MWCNTs as a function of the HNO3
concentration at 514.5 nm. The Raman intensity was normalized to that of the G band. The HNO3 concentration dependence
of the integrated intensity ratios (b) and FWHM (c) of the defect activated and G 0 Raman modes. (A color version of this figure
can be viewed online.)
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 321

hydrothermal oxidation. Such a variation indicates that be- while the excessive production of oxygenated surface groups
sides the controlled generation of functional groups, the appears to take over at higher HNO3 concentrations.
hydrothermal treatment has the capacity to reduce the To further investigate the extent of amorphous carbon
amount of amorphous/disordered carbon from the MWCNTs. generation and the thermal stability of oxygenated surface
This behavior can be contrasted to the enhanced forma- groups on the oxidized MWCNTs, the hydrothermally treated
tion of amorphous carbon during the MWCNT digestion and NTX3 and NC3100 samples at the highest HNO3 concentration
the concomitant weight loss under the harsh boiling acid oxi- (0.30 mol L1) and the boiling acid NTX3 (7.0 mol L1) samples
dation conditions [25]. This was directly evidenced by Raman were subjected to successive isothermal oxidations at 623 and
measurements on the NTX3 MWCNTs oxidized with 673 K in air. Mild oxidative treatment at 623–673 K in air is ex-
7.0 mol L1 HNO3 at boiling temperature (Fig. S4, Supplemen- pected to oxidize amorphous carbon and CCFs from CNTs
tary Information). The corresponding Raman spectra were without affecting drastically oxygen groups such as carbox-
similar to those of the pristine material rather than the ylic anhydrides and those released as CO that decompose at
0.30 mol L1 HNO3 treated sample, despite the comparable higher temperatures (Fig. 1), and thus provide an independent
amounts of functional groups produced by the boiling and probe of their modified surface chemistry upon HNO3 oxida-
hydrothermal treatments. In fact, spectral analysis revealed tion [23].
that the D, D 0, D + D 0 bands were not severely affected by Fig. 11 compares the corresponding Raman spectra of the
the boiling acid oxidation (the ID/IG, ID?/IG, ID+D?/IG intensity ra- thermally treated NTX3 samples at 514.5 nm. The Raman
tios decreased by 10% at most compared to the pristine spectra of the hydrothermally functionalized 0.30 mol L1
NTX3). On the other hand, the relative intensity of the D3 NTX3 MWCNTs showed rather small changes upon isother-
band at 1500 cm1 associated with the presence of amor- mal oxidation, the most notable variation being the decrease
phous carbon increased significantly, i.e. the ID3/IG ratio was of ID/IG by ca. 5%, which can be mainly related to the moder-
doubled for the boiling acid treated NTX3, in sharp contrast ate removal of acidic groups released as CO2 rather than
to the hydrothermally treated sample at 0.30 mol L1 HNO3, amorphous carbon [38], since the relative intensity of the D3
where the corresponding ratio showed a fivefold decrease. and D4 Raman bands was hardly affected. On the other hand,
This behavior confirmed that hydrothermal oxidation is the Raman spectra of the boiling acid treated NTX3 exhibited
much more efficient in the generation of oxygen functional- appreciable changes upon thermal oxidation demonstrating
ities on the MWCNTs compared to the aggressive boiling acid clearly the counter acting effects of hydrothermal and boiling
oxidation, where the creation of more carboxylic acid groups acid treatment on the MWCNT surface chemistry. In that
identified by TPD analysis is accompanied by extensive for-
mation of amorphous carbon on the MWCNT surface.
Fig. 10a displays the corresponding variation of the Raman
spectra for the hydrothermally functionalized NC3100 Boiling 7.0 M NTX3 514.5 nm
MWCNTs at 514.5 nm. Spectral analysis revealed that the rel- 623 K
673 K G
ative intensity of the defect activated Raman bands for the
pristine NC3100, especially that of the D, D 0 and D4 bands, D
D'
as well as their width were appreciably lower than those of G'
the corresponding NTX3 MWCNTs. In addition, the relative
intensity of the G 0 band was considerably enhanced for the
pristine NC3100 by ca. 70%, indicating that these MWCNTs
Normalized Intensity

D+D'
are more crystalline and contain smaller amounts of amor- 2G
phous carbon than the NTX3 ones, in agreement with high-
resolution TEM results on NC3100 [36]. 1000 1200 1400 1600 2400 2800 3200
Hydrothermal oxidation resulted in a moderate and most
importantly non-monotonous dependence of the ID/IG, ID 0 /IG, 0.30 M NTX3 514.5 nm
ID+D 0 /IG, ID3/IG and ID4/IG intensity ratios on the HNO3 concen- 623 K
673 K G
tration, as shown in Fig. 10b. In particular, a modest decrease G'
of the relative Raman intensity was observed up to D
0.10 mol L1 HNO3, most evident for the D band, followed by D'
a gradual increase to the initial values at higher nitric acid
concentrations. This variation was further accompanied by
the decrease of the Raman band width, especially up to D+D'
0.10 mol L1 [Fig. 10c]. Likewise, the G 0 band exhibited a de- 2G
crease of the IG 0 /IG intensity ratio together with narrowing
up to 0.10 mol L1, whereas it remained nearly stable at high- 1000 1200 1400 1600 2400 2800 3200
er HNO3 concentrations. This evolution complies favorably Raman shift (cm )
-1

with the TPD results, where a faster increment of both the


CO and CO2 amounts was observed above 0.10 mol L1 Fig. 11 – Raman spectra of the boiling (7.0 mol L1) and
HNO3. This dependence can be accordingly accounted for by hydrothermally (0.3 mol L1) acid treated NTX3 MWCNTs
the decrease of the amorphous carbon content at relatively under isothermal oxidation at 623 and 673 K. (A color
low HNO3 concentrations (0.10 mol L1), similar to NTX3, version of this figure can be viewed online.)
322 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

case, thermal oxidation led to the progressive narrowing and or changes of the D, D 0, and D + D 0 modes could be traced. The
decrease of the relative intensity of the defect activated D, D 0, lowest amount of amorphous carbon/CCFs that could be ther-
and D + D 0 modes reaching 20% at 673 K, together with an mally exfoliated from the MWCNTs’ surface was observed for
even higher reduction of the ID3/IG and ID4/IG intensity ratios the NC3100 MWCNT. This further emphasizes that apart from
(40% at 673 K) due to the decrease of the amorphous carbon the CNT morphological characteristics (diameter and length)
content. In addition, the G 0 band gradually narrowed (the the inherent structural properties and purity of the starting
FWHM decreased from 102 to 85 cm1 at 673 K) and its inten- MWCNT materials are critical factors for the efficiency of
sity ratio was greatly enhanced upon increasing the isother- the functionalization process and the nature of surface
mal annealing temperature, resembling closely the effect of groups introduced on the nanotube surface under hydrother-
hydrothermal treatment with the increase of HNO3 concen- mal oxidation.
tration on the NTX3 MWCNTs. These results confirm the ben-
eficial effects of hydrothermal HNO3 treatment that 3.4. Pore structure analysis
outperforms the harsh boiling acid oxidation by generating
controlled amounts of oxygenated groups on the CNTs, while The N2 adsorption–desorption isotherms of the pristine and
partially removing amorphous carbon fragments from the hydrothermally treated NC3100 and NTX3 samples at differ-
surface of MWCNTs. ent HNO3 concentrations are presented in Fig. 12. The
The efficiency and the mild oxidative conditions underly- NC3100 MWCNTs exhibit a typical type IV isotherm, with hys-
ing hydrothermal treatment were further corroborated by teresis at high p/p0, in agreement with previous results on
the isothermal oxidation experiments on the heavily func- NC3100 [36]. Increasing the HNO3 concentration resulted in
tionalized 0.30 mol L1 NC3100 characterized also by the high- a systematic decrease of the pore volume and widening of
est degree of graphitization. Despite the large amounts of the hysteresis loop. The total pore volume (determined from
oxygen functionalities, the Raman spectra of the thin the N2 uptake at p/p0 = 0.99) decreased from 2.9 for the pris-
NC3100 MWCNTs exhibited the least sensitivity to thermal tine NC3100 to 1.1 ml g1 for the material treated with
oxidation (Fig. S5 in Supplementary Information), where min- 0.20 mol L1 HNO3 (Table 2). The 0.30 mol L1 HNO3 treated

2000
pristine 0.05 mol L -1 0.1 mol L-1 0.2 mol L-1 0.3 mol L -1
Volume sorbed (cm3 STP g-1)

1500

1000

500

0
0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1
p/p0

1000
pristine 0.05 mol L -1 0.1 mol L-1 0.2 mol L-1 0.3 mol L -1
Volume sorbed (cm3 STP g-1)

500

0
0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1 0 0.5 1
p/p0

Fig. 12 – Adsorption–desorption N2 isotherms at 77 K of the pristine and hydrothermally functionalized NC3100 (upper row)
and NTX3 (lower row) MWCNTs at different HNO3 concentrations (left to right: 0, 0.05, 0.10, 0.20 and 0.30 mol L1 HNO3). (A
color version of this figure can be viewed online.)
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 323

Table 2 – BET surface areas, SBET, DFT micropore volume, Vmicro and pore volumes at p/p0 = 0.99, designated as Vtot (total pore
volume) and V* for the pristine and hydrothermally treated NC3100 and NTX3 MWCNTs, respectively [60].

[HNO3] (mol L1) NC3100 NTX3


2
SBET (m g ) 1 1
Vmicro (ml g ) 1
Vtot (ml g ) SBET (m2 g1) Vmicro (ml g1) V* (ml g1)

0 – pristine 278 0.004 2.9 80 0.007 0.9


0.05 375 0.022 1.8 90 0.011 1.0
0.10 409 0.029 1.4 95 0.012 1.1
0.20 429 0.034 1.1 98 0.013 1.1
0.30 441 0.039 1.1 102 0.014 1.3

sample revealed slightly higher total uptake than the 1.6


0.20 mol L1 HNO3 treated one (however similar uptake at p/ NC3100
p0 = 0.99), while the shape of the sorption isotherm at high 1.5

p/p0 pointed to the development of macroporosity (upward

Normalised S BET
1.4
turn of the isotherm at high p/p0 values). The BET surface
areas, SBET, increased continuously with the increase of the 1.3 NTX3
HNO3 concentration, as shown in Fig. 13. The values seem
to approach asymptotically a maximum correlating favorably 1.2
with the evolution of the surface group concentration on the
1.1
hydrothermally treated MWCNTs (Fig. 3).
The micropore volumes, Vmicro, remained low but also in-
1
creased with increasing HNO3 amounts (Table 2). This SBET 0 0.1 0.2 0.3 0.4
and Vmicro increase may be attributed to progressive decrease
HNO3 concentration (mol L-1)
of the amorphous carbon content as well as the enhanced
accessibility to the inner part of the nanotubes by defect cre- Fig. 13 – Normalized (with respect to pristine samples) BET
ation and/or the opening of the nanotube caps. On the other surface areas of the pristine and hydrothermally
hand, the mesopore structure of the sample is closely associ- functionalized MWCNTs at different HNO3 concentrations.
ated with the pore space between the CNTs, i.e. the bundle (A color version of this figure can be viewed online.)
free space, pointing to a more efficient arrangement of the
nanotube bundles upon hydrothermal treatment as signifi-
cant pore volume is lost. 0.09
pristine
To delineate the underlying variations in the MWCNTs
0.08 0.05 mol L-1
pore structure upon hydrothermal oxidation, pore size analy- 0.10 mol L-1
sis was performed by means of quenched solid density func- 0.07 0.20 mol L-1
0.30 mol L-1
dV(d) (cm3 STP nm-1 g-1)

tional theory (QSDFT) adsorption kernels up to a size of


0.06
50 nm. Pure cylindrical models were thereby used, while the
quenched solid variance of the method was applied as it al- 0.05
lows the study of heterogeneous surfaces [61]. The pertinent
0.04
pore size distributions (PSDs) are presented in Fig. 14. Two dif-
ferent trends were thus identified including the progressive 0.03
increase of porosity below 10 nm accompanied by the simul- 0.02
taneous decrease of the large mesopore volume upon stron-
ger treatment. Specifically, the PSDs revealed a systematic 0.01
enhancement of the small mesopore peaks (3–6 nm) centered 0.00
around 4 nm, arising from the inner part of the nanotubes, 1 2 4 6 8 10 20 40 60 100
diameter (nm)
with increasing HNO3 concentration. In this respect, it can
be concluded that hydrothermal treatment opens up more Fig. 14 – QSDFT pore size distributions of pristine and
and more pore space by e.g. creating local defects, removing hydrothermally functionalized NC3100 MWCNTs at
caps, knuckles or even debris that may originate from the different HNO3 concentrations. (A color version of this figure
CCVD synthesis. Moreover, the 0.30 mol L1 treated sample can be viewed online.)
revealed a slight decrease of the 4 nm pore volume coupled
with a stronger sign of microporosity. Such a feature may be
related to the partial deterioration of the CNT structure and
the development of amorphous (or low crystallinity) areas gether with the presence of even larger pores in the macro-
in a way that part of the inner nanotube space is narrowed. pore region. Such pores can be formed between the
On the other hand, the PSD of the pristine sample revealed entangled nanotubes, however, acid treatment leads to a con-
considerable pore volume with sizes between 6 and 50 nm, to- tinuous reduction of both their volume and size, pointing to
324 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

size decreased to ca. 32 nm for the treated samples, while


the strongest treatment (0.30 mol L1) resulted in a pore size
pristine peak at 24 nm. It may be thus concluded that the main
0.02
0.05 mol L-1 structural effect of hydrothermal treatment on the NTX3
0.10 mol L-1
dV(d) (cm3 STP nm-1 g-1)

nanotubes is a minimal disentaglement, together with a


0.20 mol L-1
0.30 mol L-1 small re-organization of the tubes (since the mean pore size
of large pores is slightly reduced).

0.01
4. Conclusions

Nitric acid hydrothermal oxidation is demonstrated as a


controlled and highly efficient functionalization process for
CCVD grown MWCNTs under mild acidic conditions. Analyt-
0.00
1 2 4 6 8 10 20 40 60 100
ical investigations by TPD–MS and TGA–DSC have shown
diameter (nm) that HNO3 hydrothermal treatment results in the progressive
generation of distinct oxygen functional groups on the
Fig. 15 – QSDFT pore size distributions of pristine and MWCNT surface, controlled by the nitric acid concentration.
hydrothermally functionalized NTX3 MWCNTs at different The amount and thermal stability of oxygen functionalities
HNO3 concentrations. (A color version of this figure can be was further found to be primarily determined by the
viewed online.) MWCNTs morphology (diameter and length), as well as by
the crystallinity and purity of the pristine materials. Further-
more, Raman spectroscopy revealed a progressive intensity
reduction along with a shift and narrowing of the defect
the nanotube disentanglement and more efficient packing
activated MWCNT Raman bands accompanied by the grad-
upon treatment. This trend can be attributed to the oxygen-
ual enhancement of the graphitic G 0 (2D) band, indicative
ated surface groups that can increase the attractive forces be-
of the oxidation of the highly reactive external carbon layers
tween individual nanotubes, e.g. through H-bonding. The
on the surface of hydrothermally treated MWCNTs with
pore sizes were centered at 33, 27, 24 and 18 nm for the pris-
increasing the HNO3 concentration. This behavior was con-
tine, 0.05, 0.10 and 0.20 mol L1 treated samples, respectively,
trasted with the enhanced amorphous carbon generation ob-
while no further significant reduction (in volume and size)
tained upon comparative processing of the CCVD MWCNTs
was observed for the NC3100 sample treated with 0.30 mol L1
with the boiling acid oxidation method that produced com-
HNO3.
parable amounts of oxygen functionalities, while consuming
On the other hand, the pore structure variations were
higher amounts of oxidizing agent.
considerably less pronounced for NTX3 MWCNTs. In that
Pore structure analysis of the functionalized MWCNTs
case, a small increase in surface area (80 up to 102 m2/g
revealed that upon hydrothermal treatment, the inner part
for the pristine and 0.30 mol L1 treated nanotubes, respec-
of the nanotubes becomes progressively more accessible
tively) and micropore volume was derived, both being signif-
to fluid molecules, leading to the increase of surface area
icantly lower than for those of the NC3100 samples (Figs. 13
(also micropore volume, which remains rather low for all
and 14 and Table 2). In contrast to NC3100, the N2 adsorption
the samples). In the case of the short and thin MWCNTs
isotherms for the NTX3 series exhibit a steep increase of
(NC3100) hydrothermal oxidation additionally triggered a
adsorption close to the saturation pressure, indicative of
disentanglement followed by more efficient reorganization
pore condensation into large meso and macropores without
of the nanotube bundles, shifting thus the mean pore size
completely filling the pore volume of the material, preclud-
from the macro to the mesopore areas. This behavior was
ing the accurate determination of the total pore volume for
considerably lessened for the long MWCNTs (NTX3), where
the NTX3 samples. Moreover, while the shapes of the iso-
the most pronounced effect of hydrothermal oxidation in-
therms remained practically the same at high p/p0, the pore
volved a small increase of the total pore volume. Nitric acid
volume values at p/p0 = 0.99 increased appreciably with
hydrothermal treatment is accordingly proposed as an
HNO3 treatment (Table 1). This behavior points to a small
effective mild functionalization process that affords the
disentanglement, however, inefficient re-packing of the
controlled modification of the MWCNTs’ surface chemistry
nanotubes during acid treatment, which can be attributed
with significant amounts of oxygen functional groups, while
to the large size (mainly length) of NTX3 MWCNTs. Further-
removing the amorphous carbon from the CCVD MWCNT
more, the corresponding QSDFT PSDs (Fig. 15) reveal rather
surface, a major drawback of conventional liquid phase oxi-
small pores at the limit of micropore sizes (around 2 nm).
dation methods.
As in the case of NC3100, these pores become more and
more accessible to fluid molecules through defects created
upon hydrothermal treatment. Beyond this size there is Acknowledgments
practically a continuum of different mesopores (e.g. local
peaks at 5 and 10 nm) indicative of the rather random nature This work was funded by the EU Seventh Framework Pro-
of the NTX3 texture in the mesoscale. Finally, much larger gramme (FP7/2007-2013) under the Clean Water project (Grant
pores (>40 nm) are observed for the pristine material; their Agreement No. 227017). Clean Water is a Collaborative Project
CARBON 6 9 (2 0 1 4) 3 1 1–32 6 325

co-funded by the Research DG of the European Commission [9] Tchoul MN, Ford WT, Lolli G, Resasco DE, Arepalli S. Effect of
within the joint RTD activities of the Environment and NMP mild nitric acid oxidation on dispersability, size, and
Thematic Priorities. This work was partially supported by pro- structure of single-walled carbon nanotubes. Chem Mater
2007;19:5765–72.
jects PTDC/AAC-AMB/122312/2010 and PEst-C/EQB/LA0020/
[10] Smith B, Wepasnick K, Schrote KE, Cho H-H, Ball WP,
2011, financed by FEDER through COMPETE and by FCT–Fun- Fairbrother DH. Influence of surface oxides on the colloidal
dacão para a Ciência e a Tecnologia. R.R.N.M. acknowledges stability of multi-walled carbon nanotubes: a structure-
FCT for her Doctoral scholarship (SFRH/BD/65425/2009). This property relationship. Langmuir 2009;25:9767–76.
research was also supported by the European Union (Euro- [11] Kim SW, Kim T, Kim YS, Choi HS, Lim HJ, Yang SJ, et al.
pean Social Fund/ESF) and Greek national funds through the Surface modifications for the effective dispersion of
carbon nanotubes in solvents and polymers. Carbon
Operational Program ‘‘Education and Lifelong Learning’’ of
2012;50:3–33.
the National Strategic Reference Framework (NSRF) – Re-
[12] Coleman JN, Khan U, Blau WJ, Gun’ko YK. Small but
search Funding Program: Thales-Investing in knowledge soci- strong: a review of the mechanical properties of carbon
ety through the European Social Fund/NANOMESO-MIS nanotube-polymer composites. Carbon 2006;44:1624–52.
377064. [13] Karousis N, Tagmatarchis N, Tasis D. Current progress on the
chemical modification of carbon nanotubes. Chem Rev
Appendix A. 2010;110:5366–97.
[14] Gerber I, Oubenali M, Bacsa R, Durand J, Gonçalves A, Pereira
MFR, et al. Theoretical and experimental studies on the
TPD spectra for pristine and hydrothermally treated NC3100
carbon-nanotube surface oxidation by nitric acid: interplay
MWCNTs at different HNO3 concentrations; TPD spectral
between functionalization and vacancy enlargement. Chem
deconvolution for NC3100 treated with HNO3 at boiling tem- Eur J 2011;17:11467–77.
perature; TGA curves for the boiling acid treated NTX3 and [15] Hu H, Zhao B, Itkis ME, Haddon RC. Nitric acid purification of
NC3100 samples; fitting analysis of the Raman spectra for single-walled carbon nanotubes. J Phys Chem B
the hydrothermally and boiling acid treated NTX3 MWCNTs 2003;107:13838–42.
in comparison with the pristine samples; Raman spectra of [16] Zhang J, Zou HL, Qing Q, Yang YL, Li QW, Liu ZF, et al. Effect
of chemical oxidation on the structure of single-walled
hydrothermally functionalized NC3100 MWCNTs after iso-
carbon nanotubes. J Phys Chem B 2003;107:3712–8.
thermal oxidation at 623 and 673 K.
[17] Yang CM, Park JS, An KH, Lim SC, Seo K, Kim B, et al.
Selective removal of metallic single-walled carbon nanotubes
Appendix B. Supplementary data with small diameters by using nitric and sulfuric acids. J Phys
Chem B 2005;109:19242–8.
Supplementary data associated with this article can be found, [18] Bergeret C, Cousseau J, Fernandez V, Mevellec J-Y, Lefrant S.
Spectroscopic evidence of carbon nanotubes’ metallic
in the online version, at http://dx.doi.org/10.1016/
character loss induced by covalent functionalization via
j.carbon.2013.12.030.
nitric acid purification. J Phys Chem C 2008;112:16411–6.
[19] Salzmann CG, Llewellyn SA, Tobias G, Ward MAH, Huh Y,
R E F E R E N C E S Green MLH. The role of carboxylated carbonaceous fragments
in the functionalization and spectroscopy of a single-walled
carbon-nanotube material. Adv Mater 2007;19:883–7.
[20] Yu H, Jin YG, Peng F, Wang HJ, Yang J. Kinetically controlled
[1] Endo M, Strano MS, Ajayan PM. Potential applications of side-wall functionalization of carbon nanotubes by nitric
carbon nanotubes. Top Appl Phys 2008;111:13–62. acid oxidation. J Phys Chem C 2008;112:6758–63.
[2] Schnorr JM, Swager TM. Emerging applications of carbon [21] Worsley KA, Kalinina I, Bekyarova E, Haddon RC.
nanotubes. Chem Mater 2011;23:646–57. Functionalization and dissolution of nitric acid treated
[3] Eklund P, Ajayan P, Blackmon R, Hart AJ, Kong J, Pradhan B, single-walled carbon nanotubes. J Am Chem Soc
et al. WTEC international assessment of research and 2009;131:18153–8.
development of carbon nanotube manufacturing and [22] Price BK, Lomeda JR, Tour JM. Aggressively oxidized ultra-
applications. Technical Report. World Technology Evaluation short single-walled carbon nanotubes having oxidized
Center Inc.; 2007. sidewalls. Chem Mater 2009;21:3917–23.
[4] Hou P-X, Liu C, Cheng H-M. Purification of carbon nanotubes. [23] Romanos GE, Likodimos V, Marques RRN, Steriotis TA,
Carbon 2008;46:2003–25. Papageorgiou SK, Faria JL, et al. Controlling and quantifying
[5] Lehman JH, Terrones M, Mansfield E, Hurst KE, Meunier V. oxygen functionalities on hydrothermally and thermally
Evaluating the characteristics of multiwall carbon nanotubes. treated single-wall carbon nanotubes. J Phys Chem C
Carbon 2011;49:2581–602. 2011;115:8534–46.
[6] Kundu S, Wang Y, Xia W, Muhler M. Thermal stability and [24] Del Canto E, Flavin K, Movia D, Navio C, Bittencourt C,
reducibility of oxygen-containing functional groups on Giordan S. Critical investigation of defect site
multiwalled carbon nanotube surfaces: a quantitative high- functionalization on single-walled carbon nanotubes. Chem
resolution XPS and TPD/TPR study. J Phys Chem C Mater 2011;23:67–74.
2008;112:16869–78. [25] Rosca ID, Watari F, Uo M, Akasaka T. Oxidation of multiwalled
[7] Wepasnick KA, Smith BA, Schrote KE, Wilson HK, carbon nanotubes by nitric acid. Carbon 2005;43:3124–31.
Diegelmann SR, Fairbrother DH. Surface and structural [26] Verdejo R, Lamoriniere S, Cottam B, Bismarck A, Shaffer M.
characterization of multi-walled carbon nanotubes following Removal of oxidation debris from multi-walled carbon
different oxidative treatments. Carbon 2011;49:24–36. nanotubes. Chem Commun 2007;5:513–5.
[8] Marques RRN, Machado BF, Faria JL, Silva AMT. Controlled [27] Datsyuk V, Kalyva M, Papagelis K, Parthenios J, Tasis D,
generation of oxygen functionalities on the surface of single- Siokou A, et al. Chemical oxidation of multiwalled carbon
walled carbon nanotubes by HNO3 hydrothermal oxidation. nanotubes. Carbon 2008;46:833–40.
Carbon 2010;48:1515–23.
326 CARBON 6 9 ( 2 0 1 4 ) 3 1 1 –3 2 6

[28] González-Guerrero AB, Mendoza E, Pellicer E, Alsina F, analysis and micro-Raman spectroscopy. J Raman Spectrosc
Fernández-Sánchez C, Lechuga LM. Discriminating the 2011;42:593–602.
carboxylic groups from the total acidic sites in oxidized [45] Singh DP, Iyer PK, Giri PK. Diameter dependence of oxidative
multi-wall carbon nanotubes by means of acid–base titration. stability in multiwalled carbon nanotubes: role of defects and
Chem Phys Lett 2008;462:256–9. effect of vacuum annealing. J Appl Phys 2010;108:084313.
[29] Solhy A, Machado BF, Beausoleil J, Kihn Y, Goncalves F, Pereira [46] Pimenta MA, Dresselhaus G, Dresselhaus MS, Cancado LG,
MFR, et al. MWCNT activation and its influence on the Jorio A, Saito R. Studying disorder in graphite-based systems
catalytic performance of Pt/MWCNT catalysts for selective by Raman spectroscopy. Phys Chem Chem Phys
hydrogenation. Carbon 2008;46:1194–207. 2007;9:1276–91.
[30] Wang Z, Shirley MD, Meikle ST, Whitby RLD, Mikhalovsky SV. [47] Delhaes P, Couzi M, Trinquecoste M, Dentzer J, Hamidou H,
The surface acidity of acid oxidized multi-walled carbon Vix-Guterl C. A comparison between Raman spectroscopy
nanotubes and the influence of in situ generated fulvic acids and surface characterizations of multiwall carbon
on their stability in aqueous dispersions. Carbon nanotubes. Carbon 2006;44:3005–13.
2009;47:73–9. [48] Fantini C, Pimenta MA, Strano MS. Two-phonon combination
[31] Rinaldi A, Zhang J, Frank B, Su DS, Abd Hamid SB, Schlogl R. Raman modes in covalently functionalized single-wall
Oxidative purification of carbon nanotubes and its impact on carbon nanotubes. J Phys Chem C 2008;112:13150–5.
catalytic performance in oxidative dehydrogenation [49] Cuesta A, Dhamelincourt P, Laureyns J, Martinez-Alonso A,
reactions. ChemSusChem 2010;3:254–60. Tascon JMD. Raman microprobe studies on carbon materials.
[32] Ambrosi A, Pumera M. Amorphous carbon impurities play an Carbon 1994;32:1523–32.
active role in redox processes of carbon nanotubes. J Phys [50] Jawhari T, Roid A, Casado J. Raman spectroscopic
Chem C 2011;115:25281–4. characterisation of some commercially available carbon
[33] Moraitis G, Spitalsky Z, Ravani F, Siokou A, Galiotis C. black materials. Carbon 1995;33:1561–5.
Electrochemical oxidation of multi-wall carbon nanotubes. [51] Dippel B, Jander H, Heintzenberg J. NIR FT Raman
Carbon 2011;49:2702–8. spectroscopic study of flame soot. Phys Chem Chem Phys
[34] Rinaldi A, Frank B, Su DS, Hamid SBA, Schlogl R. Facile 1999;1:4707–12.
removal of amorphous carbon from carbon nanotubes by [52] Sadezky A, Muckenhuber H, Grothe H, Niessner R, Poschl U.
sonication. Chem Mater 2011;23:926–8. Raman microspectroscopy of soot and related carbonaceous
[35] Ming J, Wu Y, Yu Y, Zhao F. Steaming multiwalled carbon materials: spectral analysis and structural information.
nanotubes via acid vapour for controllable nanoengineering Carbon 2005;43:1731–42.
and the fabrication of carbon nanoflutes. Chem Commun [53] Ivleva NP, Messerer A, Yang X, Niessner R, Poschl U. Raman
2011;47:5223–5. microspectroscopic analysis of changes in the chemical
[36] Tessonnier J, Rosenthal D, Hansen TW, Hess C, Schuster ME, structure and reactivity of soot in a diesel exhaust after
Blume R, et al. Analysis of the structure and chemical treatment model system. Environ Sci Technol 2007;41:3702–7.
properties of some commercial carbon nanostructures. [54] Vallerot JM, Bourrat X, Mouchon A, Chollon G. Quantitative
Carbon 2009;47:1779–98. structural and textural assessment of laminar pyrocarbons
[37] Kumar M, Ando YJ. Chemical vapor deposition of carbon through Raman spectroscopy, electron diffraction and few
nanotubes: a review on growth mechanism and mass other techniques. Carbon 2006;44:1833–44.
production. Nanosci Nanotechnol 2010;10:3739–58. [55] Frank B, Rinaldi A, Blume R, Schlogl R, Su DS. Oxidation
[38] Behler K, Osswald S, Ye H, Dimovski S, Gogotsi Y. Effect of stability of multiwalled carbon nanotubes for catalytic
thermal treatment on the structure of multi-walled carbon applications. Chem Mater 2010;22:4462–70.
nanotubes. J Nanopart Res 2006;8:615–25. [56] Shimada T, Sugai T, Fantini C, Souza M, Cançado LG, Jorio A,
[39] Silva AMT, Machado BF, Figueiredo JL, Faria JL. Controlling the et al. Origin of the 2450 cm1 Raman bands in HOPG, single-
surface chemistry of carbon xerogels using HNO 3 wall and double-wall carbon nanotubes. Carbon
hydrothermal oxidation. Carbon 2009;47:1670–9. 2005;43:1049–54.
[40] Figueiredo JL, Pereira MFR, Freitas MMA, Órfão JJM. [57] Thomsen C. Double resonant Raman scattering in graphite.
Modification of the surface chemistry of activated carbons. Phys Rev B 2000;61:4542–4.
Carbon 1999;37:1379–89. [58] Osswald S, Havel M, Gogotsi Y. Monitoring oxidation of
[41] Figueiredo JL, Pereira MFR, Freitas MA, Órfão JJM. multiwalled carbon nanotubes by Raman spectroscopy. J
Characterization of active sites on carbon catalysts. Ind Eng Raman Spectrosc 2007;38:728–36.
Chem Res 2007;46:4110–5. [59] Koh AL, Gidcumb E, Zhou O, Sinclair R. Observations of
[42] Ros TG, Dillen AJv, Geus JW, Koningsberger DC. Surface carbon nanotube oxidation in an aberration-corrected
oxidation of carbon nanofibres. Chem Eur J 2002;8:1151–62. environmental transmission electron microscope. ACS Nano
[43] Yang D-Q, Rochette J-F, Sacher E. Functionalization of 2013;7:2566–72.
multiwalled carbon nanotubes by mild aqueous sonication. J [60] Gregg SJ, Sing KSW. Adsorption, Surface Area and
Phys Chem B 2005;109:7788–94. Porosity. New York: Academic Press; 1982.
[44] Santangelo S, Messina G, Faggio G, Lanza M, Milone C. [61] Neimark AV, Lin Y, Ravikovitch PI, Thommes M. Quenched
Evaluation of crystalline perfection degree of multi-walled solid density functional theory and pore size analysis of
carbon nanotubes: correlations between thermal kinetic micro-mesoporous carbons. Carbon 2009;47:1617–28.

You might also like