Anda di halaman 1dari 6

Applied Surface Science 387 (2016) 822–827

Contents lists available at ScienceDirect

Applied Surface Science


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

Low-energy electron irradiation of preheated and gas-exposed


single-wall carbon nanotubes
P.A. Ecton a , J. Beatty b , G. Verbeck b , W. Lakshantha a , B. Rout a , J.M. Perez a,∗
a
Department of Physics, University of North Texas, Denton, TX 76203, United States
b
Department of Chemistry, University of North Texas, Denton, TX 76203, United States

a r t i c l e i n f o a b s t r a c t

Article history: We investigate the conditions under which electron irradiation at 2 keV of single-wall carbon nanotube
Received 28 March 2016 (SWCNT) bundles produces an increase in the Raman D peak. We find that irradiation of SWCNTs that are
Received in revised form 10 June 2016 preheated in situ at 600 ◦ C for 1 h in ultrahigh vacuum before irradiation does not result in an increase in
Accepted 18 June 2016
the D peak. Irradiation of SWCNTs that are preheated in vacuum and then exposed to air or gases results
Available online 21 June 2016
in an increase in the D peak, suggesting that adsorbates play a role in the increase in the D peak. Small
diameter SWCNTs that are not preheated or preheated and then exposed to air show a significant increase
Keywords:
in the D and G bands after irradiation. X-ray photoelectron spectroscopy shows no chemical shifts in the
Carbon nanotubes
Defects
C 1s peak of SWCNTs that have been irradiated versus SWCNTs that have not been irradiated, suggesting
Electron irradiation that chemisorption of adsorbates is not responsible for the increase in the D peak.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction modified with low-energy electron irradiation to produce desired


electrical properties. The mechanism by which low-energy elec-
Carbon nanotubes (CNTs) have been studied with great interest tron irradiation produces defects in bundled or isolated SWCNTs is
due to their unique structural, chemical, and electrical proper- not well understood. Only at energies greater than about 86 keV do
ties. Of particular interest is the modification of the structural electrons have sufficient momentum to cause knock-on collision
and electrical properties of CNTs through chemical, plasma, and damage to SWCNTs [9]. Since the energy of electrons in the afore-
low-energy electron irradiation processes. Numerous studies have mentioned studies is well below this threshold, the defects cannot
investigated the effects of low-energy electron irradiation on bun- be produced by knock-on collisions. It has also been reported that
dled [1–4] and isolated [4–6] single-wall CNTs (SWCNT) at energies the damage occurs as ever when SWCNTs are thoroughly degassed
of 20 eV to 20 keV and dosages of 1015 to 1018 e− /cm2 . These stud- in ultrahigh vacuum (UHV) before irradiation, with the report stat-
ies have found that the Raman spectrum of irradiated SWCNTs ing that the irradiation damage is caused by the irradiation itself
shows an increase in the intensity of the D peak located at around [4]. It would be of interest to know the conditions under which
1350 cm−1 . The D peak’s intensity is observed to decrease after low-energy electron irradiation of SWCNTs produces an increase in
annealing the sample at 300 ◦ C for 30 min. The increase in the D the D peak. In this paper, we investigate the effects of pre-heating
peak was attributed to defects in the SWCNTs caused by dam- SWCNT bundles in situ at temperatures of 400–700 ◦ C in UHV prior
age from the electron irradiation, and its reduction after annealing to low-energy electron irradiation. In addition, we study the irra-
was attributed to the healing of the SWCNTs. Studies of low- diation of SWCNT bundles that have been pre-heated in UHV and
energy electron irradiation of isolated SWCNTs have found that, in then exposed to air and gases. The samples are characterized using
addition to an increase in the D peak, irradiation induces a metal- Raman spectroscopy and X-ray photoelectron spectroscopy (XPS).
semiconductor or metal-insulator transition [4–8]. It has not been
conclusively shown if the transition is due to irradiation-induced
defects in the nanotubes [4–6] or substrate charging [7,8]. This find- 2. Material and methods
ing is particularly useful as it shows that isolated SWCNTs can be
The SWCNTs used in this study were manufactured by
Tubes@Rice using laser ablation and consist of nanotube bundles
with individual nanotubes having a distribution of diameters from
∗ Corresponding author. about 1.1 to 1.5 nm. We also studied purified SWCNTs manufac-
E-mail address: jperez@unt.edu (J.M. Perez). tured by Carbon Nanotechnologies, Inc. using the high-pressure

http://dx.doi.org/10.1016/j.apsusc.2016.06.111
0169-4332/© 2016 Elsevier B.V. All rights reserved.
P.A. Ecton et al. / Applied Surface Science 387 (2016) 822–827 823

carbon monoxide (HiPco) method that consist of nanotubes hav-


ing smaller diameters with a distribution from about 0.8 to 1.5 nm.
The nanotubes were mixed in a solution of dimethylformamide,
which was then ultra-sonicated. The mixture was then dropped on
a 0.5 × 1.5 cm2 conductive Si substrate and allowed to dry over a
24 h period.
Samples undergoing electron irradiation were first mounted
onto a quartz sample holder with two molybdenum clips attached
to allow for Ohmic heating and grounding of the substrates during
irradiation. The sample holder was then inserted into a preparation
chamber at a base pressure of 10−7 Torr. The preparation chamber
is connected via an all-metal through valve to a UHV system at
a base pressure of 10−10 Torr containing a heating stage used to
heat the samples and an electron gun used to irradiate the sam-
ples. The samples are transferred from the preparation chamber
to the UHV system using linear and rotary feedthroughs that can
transfer the sample onto the heating stage or lead the sample to
the electron gun. Samples undergoing pre-heating are heated to
temperatures from 400 to 700 ◦ C. The temperature of the sample
was measured using an external blackbody thermometer that had Fig. 1. Raman spectra of laser ablation grown SWCNTs. Black curve (a) is a spec-
trum of pristine SWCNTs that have not been electron irradiated. Red curve (b) is
line of sight access to the sample through a transparent viewport.
a spectrum of SWCNTs that have only been irradiated at 2.0 keV to a dosage of
The data presented are for samples undergoing irradiation using a 3.2 × 1015 e− /cm2 . Blue curve (c) is a spectrum of SWCNTs that have been irradiated
beam of electrons with energy set to 2.0 keV and a beam current of at 2.0 keV to a dosage of 3.2 × 1015 e− /cm2 , exposed to air, and then post-annealed
1.0 ␮A. The beam had a diameter of 1 mm and was rastered across in UHV at 450C for 1 h. Insets show the RBM region and the D band region. (For
the sample area at a rate of 0.5 mm/s. interpretation of the references to colour in this figure legend, the reader is referred
to the web version of this article.)
Samples prepared for gas exposure were returned after pre-
heating from the UHV system to the preparation chamber for
exposure to H2 O, H2 , O2 , N2 , or Ar gases. Before exposure, the Hove singularities, Eii , equal to the laser energy. SWCNTs have 6
samples were pre-heated at a temperature of 675C for 1 h. The different Raman active symmetry modes associated with in-plane
preparation chamber was primed by filling the chamber with the vibrations: two A1g , and two doubly degenerate E1g and E2g modes
intended gas at up to 100 times the base pressure. The gas was then [11]. Due to different force constants perpendicular (TO) and par-
pumped from the chamber until the chamber was brought back to allel (LO) to the tube axis, the TO and LO modes produce Raman
its base pressure. This procedure was repeated 10 times in order to peaks that are downshifted and upshifted, respectively, relative to
bring the initial residual gases down to 10−20 parts per million rel- the graphene G peak at about 1580 cm−1 [12]. In metallic SWCNTs,
ative to the intended gas. Gasses, with the exception of H2 O, were electron-phonon coupling results in a downshift and asymmetric
drawn from gas cylinders that tend to contain some water vapor. broadening of the G peak that is well fitted by a Breitt-Wigner-Fano
Because of this, before using the gases drawn from such containers (BWF) curve [13,14]. In this paper, down-shifted and up-shifted G
the tubing connecting the gas cylinder to the preparation cham- peaks are referred to as G− and G+ peaks, respectively; and, super-
ber was passed through a water vapor trap containing a mixture scripts “S” and “M” refer to semiconducting and metallic nanotubes,
of acetone and dry ice, which condensed the water vapor in the respectively.
trap preventing further diffusion into the chamber. Once the prim-
ing was complete and after the sample had been pre-heated, the 3. Results and discussion
sample was translated back to the preparation chamber and the
chamber filled with the selected gas. The pressure of the gas was Fig. 1 shows Raman spectra of pristine laser ablation grown
set to 1.00 Torr for 1 h. The gas was then removed from the prepa- SWCNTs before and after processing. The spectra have been offset
ration chamber to a pressure of 5 × 10−7 Torr, the through valve for clarity. The insets show close-ups of the radial breathing mode
opened and the sample translated back into the UHV chamber and (RBM) and D peaks with no offsets. In this and other figures, the
irradiated using the electron gun. plots in the insets are associated with the plots in the figure by color.
Electron irradiation was also carried out using electrons drawn Henceforth, pristine will refer to as-received SWCNTs that have not
from a hydrogen plasma. This method was used to irradiate sam- been pre-heated, irradiated or post-annealed. The amount of disor-
ples with electrons of lower energies (60–100 eV) that could not der in carbon is typically quantified by the ratio ID /IG , where ID is the
be produced with the electron gun. Samples undergoing plasma intensity of the D peak and IG is the intensity of the G peak. Since the
irradiation were inserted into a chamber at a base pressure of G band in SWCNTs consists of several peaks, we normalize the spec-
3 × 10−7 Torr. The plasma system has been described in detail else- tra in Fig. 1 so that the maximum of the G band at around 1590 cm−1 ,
where [10]. The plasma is generated by RF induction applied to a IM , has the same height. Fig. 1(a) shows a Raman spectrum of a pris-
quartz tube using a 50 W power supply at 20 MHz. Irradiation of the tine sample without any processing. Fig. 1(b) shows a spectrum
sample is achieved by the presence of a positive potential difference of a pristine sample that has been irradiated using the electron
between the sample and the chamber that draws the electrons from gun without in situ pre-heating or post-annealing. The D peak has
the plasma toward the sample and repels the positive ions. nearly twice the intensity as its original pristine state (ID /IM = 0.045
Raman spectra were obtained using a Nicolet Almega XR dis- and ID /IM = 0.08 for pristine and irradiated, respectively). After the
persive Raman spectrometer with a 532 nm (2.33 eV) laser. XPS Raman spectrum of the irradiated sample was taken, the sample
measurements were made using a PHI 5000 Versaprobe Scanning was post-annealed at 450C for 1 h. Fig. 1(c) shows a subsequently
XPS Microprobe. Due to their one-dimensional structure, SWCNTs acquired Raman spectrum revealing that the D peak was reduced to
have an electronic density of states that has sharp van Hove singu- just above the unprocessed level (ID /IM = 0.048), in agreement with
larities [11]. At a given laser energy, this results in resonant Raman earlier studies [5,4]. Decomposition of the RBM curve reveals three
scattering for nanotubes that have energy spacing between van Lorentzian peaks at 177, 190 and 193 cm−1 for Fig. 1(a), 177, 191
824 P.A. Ecton et al. / Applied Surface Science 387 (2016) 822–827

Fig. 3. Raman spectra of laser ablation grown SWCNTs. Black curve (a) is a spectrum
of a pristine unprocessed sample. Red curve (b) is a spectrum of a sample that was
pre-heated and irradiated in situ. Blue curve (c) is a spectrum of a sample that was
pre-heated, irradiated in situ, exposed to air and then re-irradiated. Insets show the
RBM region and the D peak region. (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)
Fig. 2. Deconvolution of the Raman curves in Fig. 1.

and 194 cm−1 for Fig. 1(b) and 173, 187 and 191 cm−1 for Fig. 1(c).
The RBM peaks correspond to the E33 S transition of semiconduct-

ing SWCNTs, as determined using a plot of Eii versus tube diameter,


dt , for a laser energy of 2.33 eV [11,15]. A small RBM component
due to metallic SWCNTs (E11 M transition) that is not observable at

the scale of Fig. 1 occurs at about 270 cm−1 . The inset also shows
the diameter of the nanotubes versus wavenumber calculated using
dt = 10 cm−1 + 234/ωRBM [16].
Fig. 2 shows an expanded view of the D and G bands in Fig. 1.
The D band is fitted with five Lorentzian peaks for each curve at
1301, 1330, 1343, 1372 and 1401 cm−1 for the pristine sample,
1310, 1332, 1344, 1373 and 1406 cm−1 for the irradiated sample,
and 1310, 1333, 1346, 1372 and 1413 cm−1 for the post-annealed
sample. In the literature, peaks at about 1310 and 1400 cm−1 have
been attributed to C H complexes [17] and asymmetric bend-
ing of CH3 [18], respectively. Peaks from 1330 to 1370 cm−1 have
been attributed to amorphous carbon, vacancies, open ends, and Fig. 4. Raman spectra of the D peaks of laser ablation grown SWCNTs exposed to
chemisorbed species such as oxides [17,19]. As shown in Fig. 2 (b), various gases after pre-heating at 675 ◦ C for 1 h. Samples include a) a pristine sample,
the most significant change after irradiation is the increase in the b) a sample undergoing pre-heating and irradiation only, samples exposed to c) O2 ,
peak at about 1333 cm−1 . The G band is fitted with four Lorentzian d) Ar, e) H2 O, f) N2 , g) H2 , and h) irradiated only SWCNTS. The inset shows the same
peaks overlapping each other to illustrate changes in intensities.
curves at around 1549, 1567, 1590, and 1607 cm−1 . As discussed in
Ref. [12], these correspond to the E2g TO , (A1g + E1g )TO , (A1g + E1g )LO ,
and E2g LO resonance modes of semiconducting SWCNTs, respec- ilar results were obtained with electrons having energies from 60 to
tively. In addition, there is a small BWF peak at around 1521 cm−1 100 eV using the plasma system. That is, the D peak increased after
and Lorentzian at around 1585 cm−1 that are assigned to the LO and irradiation if the sample was exposed to air, but did not increase if
TO modes of metallic SWCNTs, respectively [20]. the sample was pre-heated in vacuum and then irradiated in situ.
Fig. 3 shows the effects of pre-heating samples in situ. Fig. 3(a) Since air exposure is necessary to observe the phenomenon, we
shows a Raman spectrum of a pristine sample. Fig. 3(b) shows investigated the effects of exposure to specific gases. Fig. 4 shows
the spectrum of a pristine sample that has been pre-heated at the Raman spectra taken from laser-ablated SWCNTs undergoing
675 ◦ C for 1 h before electron irradiation that exhibits no increase electron irradiation at 2 keV after pre-heating in the UHV chamber
in the D peak. However, as shown in Fig. 3(c), after exposing this and gas exposure in the preparation chamber. The spectra have
sample to air for 30 min, re-installment in the UHV chamber and re- been normalized such that the maxima of the G band, IM , have
irradiation, the D peak increases (ID /IM = 0.033 and ID /IM = 0.065 for the same height. Curve (a) shows the D peak for pristine SWCNTs
the sample that was pre-heated and irradiated, and the sample that that have not been irradiated. Curve (b) shows the D peak for a
was exposed to air and re-irradiated, respectively). As shown in the sample that has been pre-heated at 675 ◦ C for 1 h and then irradi-
inset, the RBM peaks do not significantly change after irradiation; ated in situ at a dosage of 3.2 × 1015 e− /cm2 without gas exposure.
they shift by −3.7 cm−1 and increase in intensity by a factor of 0.15 Curves (c)–(g) show D peaks for samples that have been pre-heated
after pre-heating. Decomposition of the RBM curve shows three at 675 ◦ C for 1 h, and then irradiated with 2 keV electrons at a dosage
main peaks at 180, 190 and 195 cm−1 for Fig. 3(a), 170, 186 and of 3.2 × 1015 e− /cm2 after exposure to the following gases: (c) O2 ,
190 cm−1 for Fig. 3(b), and 171, 186 and 190 cm−1 for Fig. 3(c). Sim- (d) Ar, (e) H2 O, (f) N2 and (g) H2 . Curve (h) shows the D peak of
P.A. Ecton et al. / Applied Surface Science 387 (2016) 822–827 825

Fig. 5. Raman spectra of HiPco grown samples. Curve (a) is a spectrum of SWCNTs
that have not been processed. Curve (b) is a spectrum of SWCNTs that have been
irradiated and post-annealed in situ. Curve (c) is a spectrum of SWCNTs that have Fig. 6. RBM region of Raman spectra shown in Fig. 5 with corresponding labels
been irradiated, exposed to air, and post-annealed. Curve (d) is a spectrum of SWC- (a)–(g).
NTs that have only been pre-heated. Curve (e) is a spectrum of SWCNTs that have
been pre-heated and irradiated in situ. Curve (f) is a spectrum of SWCNTs that have
been pre-heated, irradiated in situ, exposed to air, and re-irradiated. Curve (g) is a
spectrum of SWCNTs that have only been irradiated. Inset (h) shows a magnification
of the RBM region.

a pristine sample that has been irradiated. For each of the gases,
hydrogen produces the highest D peak with ID /IM = 0.092, followed
by molecular nitrogen and water vapor at around ID /IM = 0.075 and
oxygen at 0.055, which is similar to pre-annealed/irradiated sam-
ples. The pristine sample has a D peak with around ID /IM = 0.065. A
slight increase in the D peak (ID /IM = 0.070) appears for Ar as well.
Because the profile of the D peak after oxygen exposure shows
no discernible difference from that of the sample that was pre-
heated and irradiated, this would suggest that oxygen does not
participate in the reaction producing the defect in the SWCNTs.
The D peak increases more after the sample has been pre-heated
and exposed to water vapor or molecular nitrogen suggesting that
water vapor and molecular nitrogen could be interacting agents. A
strong increase in the D peak after hydrogen exposure suggests
that hydrogen, a component of water, could be the main cata-
lyst. We note that in our previous experiments in which exfoliated
monolayer graphene was pre-heated, exposed to various gases and
Fig. 7. Decomposition of the D and G bands of the Raman curves in Fig. 5 with
irradiated in a similar manner, we found that molecular hydrogen corresponding labels (a)–(g).
and nitrogen exposure do not produce a D peak, but water vapor
exposure does [21]. The lack of appearance of a D peak in graphene
after molecular hydrogen and nitrogen exposure may be due to the irradiated. In these spectra, the G− band has significantly broadened
ease at which these gases desorb from graphene due to their weak and increased in intensity.
physisorption. SWCNTs tend to be more reactive due to their cur- Fig. 6 is a magnification of the RBM region of the curves
vature and nanotube bundles possess locations where molecular in Fig. 5 with corresponding labels (a)–(g). The region around
hydrogen and nitrogen can be trapped. 200–280 cm−1 shows six peaks for pre-heated samples that are
Fig. 5 shows Raman spectra of HiPco grown SWCNTs that have reduced or disappear for pristine samples or samples that have been
been processed under the same conditions as those of the laser abla- irradiated without heating. For SWCNTs exposed to a 532 nm laser
(2.33 eV), these peaks correspond to E11 M transitions for metallic
tion grown SWCNTs. Fig. 5(a) shows a spectrum of pristine SWCNTs
that have not been irradiated. The G− band is noticeably broadened SWCNTs. There also appears two peaks at around 175 and 185 cm−1
that would correspond to E33 S transitions for semiconducting nano-
and shifted to lower wave numbers in comparison to that of the
laser ablation grown semiconducting SWCNTs, consistent with it tubes. There are also peaks at around 290 cm−1 corresponding to
S transitions for semiconducting SWCNTs. Pristine SWCNTs show
E22
being due to metallic SWCNTs. Figs. 5(b)–(e) show spectra for SWC-
a slight increase in the E33S peaks and a decrease in the E M peaks
NTs that have been preheated and/or post-annealed as follows: (b) 11
irradiated and post-annealed in situ; (c) irradiated, exposed to air, with respect to heated samples and no change in the E22 S peaks.

and post-annealed; (d) only pre-heated; (e) pre-heated and irra- Samples that have been irradiated without pre-heating show a
diated in situ. In these spectra, the D peak is reduced from that of M region
large increase in the 175 cm−1 peak and a reduction in the E11
the pristine sample shown in Fig. 5(a), but the G− peaks are about with respect to heated samples.
the same. Fig. 5(f) shows the spectrum of a pristine sample that In Fig. 7, the curves in Fig. 5 were fitted with eight sub-curves;
was pre-heated, irradiated in situ, exposed to air, and re-irradiated, three Lorentzian curves at around 1310, 1325 and 1372 cm−1 for
and Fig. 5(g) shows the spectrum of a pristine sample that was only the D band, and four Lorentzian curves centered at around 1517,
826 P.A. Ecton et al. / Applied Surface Science 387 (2016) 822–827

Fig. 9. XPS of HiPco grown SWCNTs. Similar to the laser ablation grown SWCNTs,
Fig. 8. XPS of laser ablation grown SWCNTs. No difference is indicated between the no difference is found between the pristine and irradiated samples. A slight shift
as-prepared pristine and irradiated samples. Annealing shows a decrease in the C O of about 0.2 eV is apparent between the annealed samples and the non-annealed
peak as well as a reduction in the FWHM of the C1s peak. samples.

1542, 1572, and 1585 cm−1 , and a Breit-Wigner-Fano (BWF) curve We characterize the defects produced in SWCNT bundles by
centered at 1518 cm−1 for the G band. Peak assignments and sym- their effect on the BWF peak. The BWF peak in bundles has been
metries are extrapolated from previous studies [16]. The BWF curve reported to increase as a result of tube–tube interactions that are
is attributed to phonon modes of A1g symmetry in metallic nano- thought to produce a plasmon band that increases the plasmon-
tubes and is asymmetric due to electron phonon coupling [20]. The phonon coupling [24]. For example, when bundles are reduced in
most likely source of this curve in the samples comes from the size [24], or the nanotubes within a bundle are isolated by oxi-
metallic nanotubes with E11 M transition. The 1517 cm−1 curve can dation [25] or wrapping with DNA molecules [22,23], the BWF
be attributed to a G−S peak from the small diameter E22 S semicon- peak decreases. The BWF peak increases again when the oxide or
ducting nanotubes of E2g symmetry overlapped with a G−M peak DNA molecules are removed. In addition, defects such as vacan-
from mid diameter nanotubes with E11 M transitions and A symme- cies, oxides and cut edges produced by high power pulsed laser
1g
try. The 1542 cm−1 curve can be attributed to the G−M peak from irradiation decrease the BWF peak [17,26]. Since the BWF peak
upper mid diameter metallic nanotubes with E11 M transitions and increases as a result of an increase in tube–tube interactions, we
A1g symmetry in which inter-tube electron-phonon coupling does propose that the defects formed by low-energy electron irradia-
not appear to occur [22,23]. The 1572 cm−1 curve can be attributed tion are such that they increase tube–tube interactions. This would
to the G−S peak of the large diameter semiconducting nanotubes exclude oxides, vacancies and cut edges since they decrease the
with E33S transitions and E −1 curve BWF peak. A defect that may increase tube–tube interactions is
1g symmetry, and the 1585 cm
+
can be attributed to the G peak for both metallic and semiconduct- inter-tube linking in which adjacent tubes form bridges. Inter-tube
ing nanotubes. The lower three curves (BWF and two Lorentzians) linking has been reported to form in SWCNT bundles as a result of
that contain metallic components tend to significantly increase electron irradiation at an energy of 80 keV [27,28]. Although this
after irradiation without heating but decrease when heated. These energy is about an order of magnitude higher than those used in
curves correspond to nanotubes that have smaller diameters than our experiments, it is below the theoretical knock-on threshold of
those grown using laser ablation. Their greater curvature may make 86 keV. We note that cross-linking has been reported in C60 films
them more susceptible to damage from irradiation. irradiated with electrons at energies comparable to those used in
Since gas exposure is required for the phenomenon to occur, our experiments. Adjacent C60 molecules in such films have been
we carried out XPS measurements to determine if physisorbed observed to form cross-linked peanut-shaped structures as a result
gases chemisorbed on the nanotubes as a result of the irradiation. of irradiation at 3 keV [29]. Post-annealing at 200 ◦ C for 2 h reverts
Fig. 8 shows the XPS spectra of the C 1s region of four laser abla- the peanut-shaped structures back to C60 [30].
tion grown SWCNT samples; a pristine sample, a pristine sample
that was irradiated without pre-heating, a pristine sample that was 4. Conclusions
irradiated and then post-annealed in situ (600C for 30 min), and a
pre-heated sample (600 ◦ C for 30 min) that was irradiated in situ. Our work shows that low-energy electron irradiation of SWCNT
No noticeable change between the spectra of a pristine sample bundles itself does not cause damage; we find that, for damage to
and an irradiated pristine sample is observed. Both spectra show occur, there must also be adsorbates present on the nanotubes. To
peaks due to C O bonds and no peaks due to hydrogen or nitrogen our knowledge, this has not been reported to date. This conclu-
species. The XPS spectra of the two samples that underwent pre- sion is based on our observation that the D peak does not increase
heating and post-annealing are similar and show that the oxygen after adsorbates have been removed from the sample through the
species that were originally chemisorbed on the pristine samples process of preheating in UHV at 600 ◦ C before electron irradiation
have desorbed. Fig. 9 shows XPS spectra of HiPco grown SWCNTs is performed, but does increase after the sample is preheated and
that were similarly processed showing again no noticeable change exposed to air or gases before irradiation. We find that the influence
between the spectra of a pristine sample and an irradiated pris- of gas adsorbates is more significant on the HiPco-grown samples
tine sample. Since the XPS data show no change in the chemical than the laser-ablation grown samples. The laser-ablation grown
samples consisted mainly of large-diameter E33 S semiconducting
composition of the CNTs, this indicates that although the adsorbed
gases mediate the defect formation, they do not strongly chemisorb. SWCNTs, while the HiPco-grown samples consisted of E11 M metallic

Therefore, the adsorbed gases may facilitate the production of a SWCNTs that have smaller diameters than the laser-ablation grown
structural defect in the C lattice and remain physisorbed, or pro- S semiconducting SWCNTs that have smaller
samples, as well as E22
duce a charge-transfer complex that distorts the lattice but does M
diameters than the E11 metallic SWCNTs. It is known that the Fermi
not strongly chemisorb. level electrons in metallic SWCNTs can participate in chemical reac-
P.A. Ecton et al. / Applied Surface Science 387 (2016) 822–827 827

tions. Such electrons are absent in semiconducting SWCNTs. It is [12] A. Jorio, G. Dresselhaus, M.S. Dresselhaus, M. Souza, M.S.S. Dantas, M.A.
also known that as the diameter of a SWCNT decreases, the reac- Pimenta, A.M. Rao, R. Saito, C. Liu, H.M. Cheng, Polarized Raman study of
single-wall semiconducting carbon nanotubes, Phys. Rev. Lett. 85 (2000) 2617.
tivity increases due to the increase in curvature and resulting bond [13] M. Lazzeri, S. Piscanec, F. Mauri, A.C. Ferrari, J. Robertson, Phonon linewidths
strain [31]. We attribute the greater influence of adsorbates on and electron-phonon coupling in graphite and nanotubes, Phys. Rev. B 73
the HiPco-grown samples to the Fermi level electrons and smaller (2006) 155426.
M metallic SWCNTs, and the smaller diameters [14] S. Piscanec, M. Lazzeri, J. Robertson, A.C. Ferrari, F. Mauri, Optical phonons in
diameters of the E11 carbon nanotubes: Kohn anomalies, Peierls distortions, and dynamic effects,
S semiconducting SWCNTs. Both laser-ablation and HiPco
of the E22 Phys. Rev. B 75 (2007) 035427.
grown samples recover after post-annealing at 450 ◦ C. XPS spectra [15] H. Kataura, Y. Kumazawa, Y. Maniwa, I. Umezu, S. Suzuki, Y. Ohtsuka, Y.
Achiba, Optical properties of single-wall carbon nanotubes, Synth. Met. 103
shows no evidence for irradiation-induced chemisorption as the (1999) 2555–2558.
cause for the increase in the D peak. Since the BWF peak in metallic [16] M.S. Dresselhaus, G. Dresselhaus, R. Saito, A. Jorio, Raman spectroscopy of
SWCNTs increases after irradiation, we propose that the increase in carbon nanotubes, Phys. Rep. 409.2 (2005) 47–99.
[17] M. Tachibana, Characterization of Laser-Induced Defects and Modification in
D peak for such nanotubes is due to defects that increase tube–tube Carbon Nanotubes by Raman Spectroscopy, INTECH Open Access Publisher,
interactions, such as inter-tube linking. Since isolated SWCNTs have 2013.
been reported to show an increase in the D peak after low-energy [18] M. Shahzad Imran, M. Giorcelli, N. Shahzad, S. Guastella, M. Castellino, P.
Jagdale, A. Tagliaferro, Study of carbon nanotubes based
electron irradiation, it would be interesting to investigate if pre- polydimethylsiloxane composite films, J. Phys.: Conf. Ser. 439 (2013) 012010.
heating isolated SWCNTs in UHV before irradiation prevents an [19] A.C. Ferrari, J. Robertson, Resonant Raman spectroscopy of disordered,
increase in the D peak and occurrence of a metal-semiconductor amorphous, and diamondlike carbon, Phys. Rev. B 64 (7) (2001) 075414.
[20] S.D.M. Brown, A. Jorio, P. Corio, M.S. Dresselhaus, G. Dresselhaus, R. Saito,
transition.
et al., Origin of the Breit-Wigner-Fano lineshape of the tangential G-band
feature of metallic carbon nanotubes, Phys. Rev. B 63 (2001) 155414.
References [21] J.D. Jones, K.K. Mahajan, W.H. Williams, P.A. Ecton, Y. Mo, J.M. Perez,
Formation of graphane and partially hydrogenated graphene by electron
irradiation of adsorbates on graphene, Carbon 48 (2010) 2335–2340.
[1] S. Suzuki, K. Kanzaki, Y. Homma, S.Y. Fukuba, Low-acceleration-voltage
[22] S.G. Chou, H.B. Ribeiro, E.B. Barros, A.P. Santos, D. Nezich, G.G. Samsonidze, C.
electron irradiation damage in single-walled carbon nanotubes, Jpn. J. Appl.
Fantini, M.A. Pimenta, A. Jorio, F. Plentz Filho, M.S. Dresselhaus, Optical
Phys. 43 (8B) (2004) L1118.
characterization of DNA-wrapped carbon nanotube hybrids, Chem. Phys. Lett.
[2] S. Suzuki, Y. Kobayashi, Diameter dependence of low-energy electron and
397 (4) (2004) 296–301.
photon irradiation damage in single-walled carbon nanotubes, Chem. Phys.
[23] H. Kawamoto, T. Uchida, K. Kojima, M. Tachibana, The feature of the
Lett. 430 (4) (2006) 370–374.
Breit-Wigner-Fano Raman line in DNA-wrapped single-wall carbon
[3] S. Suzuki, K. Yamaya, Y. Homma, Y. Kobayashi, Activation energy of healing of
nanotubes, J. Appl. Phys. 99 (9) (2006) 094309.
low-energy irradiation-induced defects in single-wall carbon nanotubes,
[24] C. Jiang, K. Kempa, J. Zhao, U. Schlecht, U. Kolb, T. Basché, M. Burghard, A.
Carbon 48 (2010) 3211–3217.
Mews, Strong enhancement of the Breit-Wigner-Fano Raman line in carbon
[4] S. Suzuki, in: J.M. Marulanda (Ed.), Low-Energy Irradiation Damage in
nanotube bundles caused by plasmon band formation, Phys. Rev. B 66 (16)
Single-Walled Carbon Nanotubes, Electronic Properties of Carbon Nanotubes,
(2002) 161404.
InTech, 2011 (ISBN: 978-953-307-499-3).
[25] Z. Yu, L. Brus, Rayleigh and Raman scattering from individual carbon
[5] K. Kanzaki, S. Suzuki, H. Inokawa, Y. Ono, A. Vijayaraghavan, Y. Kobayashi,
nanotube bundles, J. Phys. Chem. B 105 (6) (2001) 1123–1134.
Mechanism of metal-semiconductor transition in electric properties of
[26] D. Kang, K. Kato, K. Kojima, M. Tachibana, T. Uchida, Phonon control in
single-walled carbon nanotubes induced by low-energy electron irradiation, J.
metallic carbon nanotubes due to laser-induced defects, App. Phys. Lett. 93
Appl. Phys. 101 (2007) 034317.
(2008) 133102.
[6] S. Suzuki, Y. Kobayashi, Low-energy irradiation damage in single-walled
[27] A. Kis, G. Csanyi, J.P. Salvetat, T.N. Lee, E. Couteau, A.J. Kulik, W. Benoit, J.
carbon nanotubes, Cambridge University Press, in: MRS Proceedings, Vol. 994,
Brugger, L. Forro, Reinforcement of single-walled carbon nanotube bundles by
2007, pp. 0994-F04.
intertube bridging, Nat. Mater. 3 (3) (2004) 153–157.
[7] A. Vijayaraghavan, K. Kanzaki, S. Suzuki, Y. Kobayashi, H. Inokawa, Y. Ono,
[28] P. Jaroenapibal, D.E. Luzzi, S. Evoy, Tuning the resonant frequency of
Metal-semiconductor transition in single-walled carbon nanotubes induced
single-walled carbon nanotube bundle oscillators through
by low-energy electron irradiation, Nano Lett. 5.8 (2005) 1575–1579.
electron-beam-induced cross-link formations, Appl. Phys. Lett. 90 (8) (2007)
[8] C.W. Marquardt, S. Dehm, A. Vijayaraghavan, S. Blatt, F. Hennrich, R. Krupke,
081912.
Reversible metal-insulator transitions in metallic single-walled carbon
[29] J. Onoe, T. Nakayama, M. Aono, T. Hara, Structural and electrical properties of
nanotubes, Nano Lett. 8 (2008) 2767–2772.
an electron-beam-irradiated C60 film, Appl. Phys. Lett. 82 (4) (2003) 595–597.
[9] F. Banhart, Irradiation effects in carbon nanostructures, Rep. Prog. Phys. 62
[30] Y.B. Zhao, D.M. Poirier, R.J. Pechman, J.H. Weaver, Electron stimulated
(1999) 1181.
polymerization of solid C60, Appl. Phys. Lett. 64 (5) (1994) 577–579.
[10] J.D. Jones, W.D. Hoffman, A.V. Jesseph, C.J. Morris, G.F. Verbeck, J.M. Perez, On
[31] S. Niyogi, M.A. Hamon, H. Hu, B. Zhao, P. Bhowmik, R. Sen, M.E. Itkis, R.C.
the mechanism for plasma hydrogenation of graphene, Appl. Phys. Lett. 97
Haddon, Chemistry of single-walled carbon nanotubes, Acc. Chem. Res. 35
(2010) 233104.
(12) (2002) 1105–1113.
[11] R. Saito, G. Dresselhaus, M.S. Dresselhaus, Physical Properties of Carbon
Nanotubes, vol. 35, Imperial College Press, London, 1998.

Anda mungkin juga menyukai