Vous êtes sur la page 1sur 6

Carbon 95 (2015) 28e33

Contents lists available at ScienceDirect

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

Controlling the thickness of carbon nanotube random network lms


by the estimation of the absorption coefcient
Francesco De Nicola a, b, *, Chiara Pintossi c, Francesca Nanni d, Ilaria Cacciotti e,
Manuela Scarselli a, b, Giovanni Drera c, Stefania Pagliara c, Luigi Sangaletti c,
Maurizio De Crescenzi a, b, Paola Castrucci a, b
 di Roma Tor Vergata, Via della Ricerca Scientica 1, 00133 Roma, Italy
Dipartimento di Fisica, Universita
 di Roma Tor Vergata (INFN-Roma Tor Vergata), Via della Ricerca Scientica 1, 00133 Roma, Italy
Istituto Nazionale di Fisica Nucleare, Universita
I-LAMP and Dipartimento di Matematica e Fisica, Universit Cattolica del Sacro Cuore, Brescia, Italy
d
 di Roma Tor Vergata (INSTM-UdR Roma Tor Vergata), Via del Politecnico 1, 00133 Roma, Italy
Dipartimento di Ingegneria dell'Impresa, Universita
e
 di Roma Niccolo
 Cusano (INSTM-UdR), Via Don Carlo Gnocchi 3, 00166 Roma, Italy
Universita
a

b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 23 March 2015
Received in revised form
30 July 2015
Accepted 31 July 2015
Available online 7 August 2015

Here, we investigate the thickness of single-walled (SWCNT) and multi-walled carbon nanotube
(MWCNT) random network lms by angle-resolved X-ray photoemission spectroscopy. Furthermore, we
estimate the absorption coefcient of carbon nanotube (CNT) lms through the LamberteBeer law, by
measuring lm optical spectra. Moreover, the knowledge of the absorption coefcient provides an easier,
reliable, and faster method of investigation for generic CNT lm thickness. In addition, the absorption
coefcient leads to the information of the absorption length for SWCNT and MWCNT lms, which is a
physical quantity of fundamental interest for optoelectronic applications, such as light emitting diodes,
photovoltaics, and in general light absorbers.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
The thickness of carbon nanotube (CNT) lms is a fundamental
quantity in order to understand their optical and electronic properties. However, measurements of the lm thickness are often
fraught with large and undesirable uncertainties, owing to the thin
lm high degree of roughness or in general to its complex
morphology [1]. In particular, this is notably for non-vertically
aligned CNT lms. Therefore, local scanning probe microscopy
such as atomic force microscopy (AFM) and even non-local, such as
scanning electron microscopy (SEM) cannot provide a reliable
estimation of the CNT lm thickness, because they usually are very
sensitive to the CNT lm surface roughness [2].
A closely related quantity to the CNT lm thickness is its absorption or extinction coefcient. The latter is of paramount
importance to characterize the optical and optoelectronic properties of CNT lms [3]. So far, the estimation of CNT absorption

 di Roma Tor Vergata,


* Corresponding author. Dipartimento di Fisica, Universita
Via della Ricerca Scientica 1, 00133 Roma, Italy.
E-mail address: fdenicola@roma2.infn.it (F. De Nicola).
http://dx.doi.org/10.1016/j.carbon.2015.07.096
0008-6223/ 2015 Elsevier Ltd. All rights reserved.

coefcient has been reported for CNT dispersions [4,5] by means of


the molar extinction coefcient, and only in the infrared range for a
vertically aligned CNT solid lm [6]. Nevertheless, there is no reported information about CNT random network lm absorption
coefcient. Similarly, the knowledge of the CNT lm absorption
length or penetration depth may result crucial for CNT-based optical and optoelectronic applications. Moreover, the CNT lm
thickness estimation by the absorption coefcient (i.e. by spectrophotometry) is a reliable, easier, and faster investigation method,
which provide an information of the rough material effective
thickness.
In the following, vacuum-ltered single-walled carbon nanotube (SWCNT) and multi-walled carbon nanotube (MWCNT) lms
are realized, and the lm thicknesses is estimated by angleresolved X-ray photoemission spectroscopy (AR-XPS), which has
proven to be excellent tool for non-destructive depth proling of
nanostructured materials and interfaces [7e9]. In addition, the
absorption coefcients and lengths of SWCNT and MWCNT lms
are obtained over the UV/Vis/NIR range. We show that CNT lms
are exceptional and broadband absorbers and may be used not only
for photovoltaic applications [10e17] and as bolometers [18e26],
but also as light emitters [27e30]. Moreover, we demonstrate that

F. De Nicola et al. / Carbon 95 (2015) 28e33

29

our method, not only may be very useful and effective for the
estimation of generic CNT lm thickness, but also can be used to
tailor in a controlled fashion the thickness of CNT lms, simply on
the basis of the volume of the CNT dispersion.
2. Experimental details
2.1. Fabrication of CNT lms
Highly pure CoMoCAT SWCNT powder (SigmaeAldrich, assay
>90%, diameter: 0.7e0.9 nm) and CCVD MWCNT powder (Nanocyl,
NC3100, assay >90%, diameter: 5e9 nm) were dispersed with
several concentrations in aqueous solution with 2% w/v sodiumdodecil-sulfate (SigmaeAldrich, assay >98.5%) anionic surfactant.
In addition, to better disperse the suspension, CNTs were tipultrasonicated (Branson S250A, 200 W, 20% power, 20 KHz) in an
ice-bath for an hour and the unbundled supernatant was collected
by pipette. The result was a well-dispersed suspension which is
stable for several months. Carbon nanotube lms were fabricated
by a vacuum-ltration process of aliquot volumes of the dispersion
cast on mixed cellulose ester lters (Pall GN6, 1 in diameter,
0.45 mm pore diameter). Subsequently, rinsing in water and in a
mixture of ethanol, methanol, and water (15:15:70) to remove as
much surfactant as possible was performed. Samples were made
depositing by the dry-transfer printing method [1,2] two cut parts
of the same CNT lm; one on Carlo Erba soda-lime glass slides for
optical absorption measurements, and the other part on HF etched
silicon (100) wafers for AR-XPS experiments.
2.2. Optical spectroscopy and electron microscopy
Optical spectroscopy (Perkin-Elmer Lambda 19 UV/Vis/NIR) was
performed to characterize the absorption coefcient of CNT lms.
We measured the transmittance T(l) I/I0 spectra of CNT lms for
wavelengths l in the UV/Vis/NIR range, where I and I0 are the intensities of light transmitted by the CNT lm deposited on a glass
substrate and by a reference bare glass substrate, respectively. We
then calculated the corresponding absorbance A(l) spectra by the
known equation A(l) log10T(l). Scanning electron microscopy
micrographs were acquired with Zeiss Leo Supra 35 eld emission
scanning electron microscope (FEG-SEM).
2.3. AR-XPS measurements and modeling
The CNT lm thickness was measured by AR-XPS measurements
of CNT lms deposited on silicon wafers, as previously reported [8].
The AR-XPS data were collected with a SCIENTA R3000 analyzer,
operating in transmission mode and working with an acceptance
angle of 30 . Based on the electrostatic lens setting of the analyser
and the X-ray source spot on the target, the sampled area is about
3  3 mm2. The AlKa line (hn 1486.6 eV, resolution 0.85 eV) of a
non-monochromatized dual-anode PsP X-ray source was
employed, running at a power of about 110 W. The base pressure in
the sample analysis chamber was z1010 mbar. With the sample
holder clips being coated with gold, the binding energy
(Eb 83.96 eV) [31] of the Au 4f7/2 peak was taken as a reference.
The measurement of the CNT layer thickness is based on the
attenuation of the core level intensity signal from the Si substrate
for different CNT coverages. This measurement can be carried out at
different extents, from the attenuation determined by inelastic
scatting events alone, to more sophisticated models that also
include elastic scattering processes and the use of the depth distribution function (DDF). In the present case, the CNT layer thickness was estimated by collecting the C 1s and Si 2p core level
intensities at different take-off angles q (Fig. 1).

Fig. 1. Sketch of AR-XPS measurements for CNT lms deposited on silicon.

The thickness of each layer in the model was estimated by


considering the samples as a CNT-Si interface. Modeling [32e35]
and simulation of the multilayer structure in order to t the ARXPS peak angular dependence was carried out by the BriXias
code [36] developed at the Surface Science and Spectroscopy Lab
and based on the IGOR Pro 6.2 programming language. A variety of
physical processes are known to play an important role in the ARXPS signal generation, such as elastic and inelastic scattering in
the interior of the solid, surface excitations, and intrinsic excitations
following the ionization that precedes the signal electron emission
in AR-XPS. Generally, such processes can occur repeatedly (multiple
scattering) before the electron is ejected from the surface, giving
rise to a complex combination of effects responsible for some
important features observed in experimental spectra. The AR-XPS
core level peak intensities I(Ek,q) of a selected layer at a depth
d with a thickness t can be evaluated through the formula [35,36].

Zdt
IEk ; q K

FEk ; q; zdz;

(1)

where K is a normalization constant, which includes the photoionization cross section, the atomic density of the species, and
analyzer-dependent parameters; F(Ek,q,z) is the generic escape
probability or DDF of an electron generated at a depth z with a
kinetic energy Ek at an angle q with respect to the surface normal.
According to the LamberteBeer law, the DDF function is usually
approximated with a Poisson distribution [33] F exp(z/Lcosq)
where L is the inelastic mean free path of the photoelectrons [37].
Such an approximation leads to a simple analytical expression for
the peak areas, which is dened as a straight line motion [35].
However, this formulation may result in an overestimation of the
top layer thickness, especially for thin overlayers. Here, we used
DDF calculations [33], in order to account for both inelastic and
elastic electronic scattering processes in the so-called transport
approximation [34]. Furthermore, the photoemission asymmetry
parameters are taken into account for each core level and calculations of electronic trajectories are computed to predict the AR-XPS
peak areas in the CNT-Si interface, since an analytic DDF formulation [35] cannot be written for a generic multilayer sample.
3. Results and discussion
3.1. Measure of the CNT lm thickness
Several samples were realized by ltering different volume aliquots of the same CNT aqueous dispersion. The obtained lms are
random networks of SWCNTs or MWCNTs, as evident from SEM
micrographs in Fig. 2aed. From grazing incidence SEM images in

30

F. De Nicola et al. / Carbon 95 (2015) 28e33

Fig. 2c,d, it can be observed that the high roughness due to the
complex microstructure of CNT lms makes the estimation of the
lm thickness a hard task. Therefore, in order to overcome this
issue, we performed AR-XPS measurements on CNT lms, which
provides an estimation of the average lm thickness probed by
electrons. The overall trend of the angular dependence of the
photoemission spectra is detectable in Fig. 3a, which shows the ARXPS data of C 1s and Si 2p core levels collected for a MWCNT-Si
interface at different take-off angles q. These data have been
collected as survey spectra with a pass energy of the analyzer set at
100 eV and averaging over 50 scan for each spectrum. As can be
observed, both the C 1s and Si 2p intensities change with the takeoff angle. In particular, as q increases, the measure becomes more
surface sensitive, therefore the signal intensity from the Si substrate decreases, while the intensity of the C signal increases. For a
detailed data analysis and modeling, high resolution and low-noise
AR-XPS spectra have subsequently been collected by setting the
analyzer pass energy at 50 eV and averaging the signal over 1300
scans for each spectrum. In order to evaluate the CNT lm average
thickness, we used the model described in Section 2.3, where the
intensity of each AR-XPS peak is strongly dependent on the thickness t of the layer itself and on the thickness d of all the other layers
above (see equation (1)). This means that the AR-XPS intensity of
any layer under the CNT lm is connected to the thickness of the
CNT lm itself. In this framework, it is fundamental to proceed with
a suitable normalization (involving both sides of equation (1)) of
the AR-XPS peak intensity, to properly account for instrumental
factors which are independent on the take-off angle and on the
sample itself. Therefore, the Si 2p spectra have been normalized to
the C 1s intensity. The sequence of Si 2p AR-XPS spectra, normalized to the C 1s peak area and collected at different take-off angles,
is shown in Fig. 3b. Following Ref. [8] the Si 2p were tted by three
or four Voigt peaks accounting for bulk Si (one peak), SiO2 (one
peak), and a non-stoichiometric silicon oxide (SiOx) layer which
develops at the CNT-Si interface, as the samples are stored in air. In
particular, one and two peaks were used for the SiOx contribution in
MWCNT and SWCNT respectively, to account for the different
extent of the oxidation processes at these two CNT-Si interfaces.
The binding energies of the Si related peaks are: 103.4 eV (SiO2),
100.1 eV (SiOx), 99.5 eV (Si). The origin of the peaks are discussed in

Ref. [38,39]. The angular dependence of the intensities of the


relevant AR-XPS peaks is shown in Fig. 3c (dots) along with the
corresponding ts (solid lines). The tted thickness values (i.e. the
free parameters in the model) yielding our results were assumed as
the best estimate of the actual thickness of each layer in the model.
This method was applied to all samples considered in the present
study and their thickness values are reported in Table 1. We
assumed in our model a bulk density of the CNT lms r 0.128 g/
cm3 (SigmaeAldrich). Such a CNT low density allowed us to probe
by AR-XPS not only a relative thick (25e40 nm) CNT lm but also
the buried Si substrate underneath. Indeed, for random network
CNT lms the electrons emitted at a certain depth may cross, in
their way to the detector, both CNT bundles and voids. Assuming
that voids do not perturb the electron trajectory (both in terms of
electron kinetic energy and momentum) the electron mean free
path is longer than the usual 2e3 nm. Moreover, it is worth noting
that within our model the CNT lm roughness can be accounted for
by considering the emission from CNT layers with different heights
randomly distributed on the sample surface, which ultimately
result in an average thickness of the CNT layer. In this perspective,
the resulting thickness of the CNT layer can be regarded as an
effective value.
3.2. Estimation of the absorption coefcient and absorption length
of CNT lms
Optical spectroscopy was performed to estimate the absorption
coefcient of several CNT lms. In Fig. 4a, we report as example the
absorption spectra of few SWCNT and MWCNT lms in the UV/Vis/
NIR range. It is clear from the plot that for a given CNT dispersion
with its own concentration the optical absorption increases with
the volume aliquots used to make the CNT lm. Also, we may relate
absorption peaks in our SWCNT lm spectra to different optical
transitions of semiconducting (Sii, z90% in percentage) and
metallic (M11, z10%) SWCNTs [15,40]. Therefore, we infer that the
realized SWCNT lms consist of several electronic type or chirality,
in the range of nanotube diameters z0.7e0.9 nm, with relevant
semiconducting and metallic features. On the other hand, we did
not observe any relevant feature in MWCNT lm spectra, although
the spectra are very similar to those of SWCNT lms. The pp*

Fig. 2. Scanning electron micrographs of SWCNT (a,c) and MWCNT (b,d) lms at different magnications 10,000 (c,d) and 200,000 (a,b). Carbon nanotube lms appear as dense
and porous (dark holes) random networks (a,b). In the image taken at grazing incidence (c,d) (z90 respect with the plane normal), it is possible to observe micro-structures
consisting in self-assembly ripples made of SWCNTs. Conversely, MWCNTs just aligned in the vertical direction. Dark areas is the underneath glass substrate.

F. De Nicola et al. / Carbon 95 (2015) 28e33

31

Table 1
Results of AR-XPS and spectrophotometry measurements.
Sample

Film thickness (nm)

SWCNT01
SWCNT02
SWCNT03
MWCNT01
MWCNT02
MWCNT03

29.2
31.4
36.1
22.2
28.6
35.4

0.8
0.9
2.2
0.8
1.2
5.3

Absorbance at 550 nm (o. d.)


0.11
0.24
0.45
0.14
0.63
1.00

surface plasmon polariton absorption peak (Fig. 4a), which is a


typical feature of graphitic materials [41e48], is indeed present in
both CNT lm spectra at 253 nm (5.1 eV) and at 243 nm (4.9 eV) for
the MWCNT and SWCNT lms, respectively. Moreover, the CNT lm
absorbance A(l) at a given wavelength l is linearly dependent on
the lm thickness t and on the absorption coefcient a through the
LamberteBeer law

Al alt:

(2)

We plotted for SWCNT and MWCNT lms their thickness values


estimated by AR-XPS as a function of the absorbance measured for
instance at 550 nm (Fig. 4b). Data are also reported in Table 1. Then,
by linearly tting the experimental data, we found for the SWCNT
lm aSWCNT(550 nm) (5.3 0.6)$105 cm1 and for the MWCNT
lm aMWCNT(550 nm) (7.4 0.5)$105 cm1. However, ts return a
non-zero intercept b, which can be attributed to residual SDS surfactant absorption [49]. These lm thickness independent absorptions are respectively for our SWCNT and MWCNT lms
bSWCNT 1.4 0.1 o. d. and bMWCNT 1.5 0.1 o. d. Our CNT absorption coefcient values are comparable with the value calculated for bulk graphite (4$105 cm1 [50]) and the value obtained by
the dielectric function measured by electron energy loss spectroscopy (EELS) on individual SWCNTs (2$105 cm1 [46]) and by
ellipsometry on aligned MWCNT lms (2$105 cm1 [51]). In Fig. 4c
the two CNT absorption coefcient spectra in the UV/Vis/NIR range
are reported. Moreover, in Fig. 4d we can observe the CNT absorption length d a1. For instance, it is necessary only a 24 nm
thick SWCNT lm or a 14 nm MWCNT lm thick to absorb all the
light within 1000 nm in wavelength. Therefore, CNT lms can be
exceptional and broadband absorbers. It is worth noting that, as
shown in Fig. 5, the absorption coefcients found for our SWCNT
and MWCNT random network lms are well above those of conventional semiconductors used in actual solar cell devices and light
emitting diodes, such as crystalline Si, amorphous Si (a-Si), GaAs,
AlGaAs, CdTe, CuInSe2, CdS, Ge, and InP.
3.3. Calibration of the CNT lm thickness

Fig. 3. (a) Angle-resolved X-ray photoemission survey scans of the C 1s and Si 2p core
level regions for the sample MWCNT01 deposited on Si at different take-off angles
q 0 ,10 ,20 ,30 ,40 ,50 . (b) High resolution Si 2p AR-XPS dataset. All the Si 2p peaks
were normalized to the C 1s and tted with three Voigt curves ascribed to SiO2 (peak 1
in red), SiOx (peak 2 in gray) and bulk Si (peak 3 in blue). (c) SiO2 (red dots), SiOx (gray
dots), and Si (blue dots) peak areas as functions of take-off angles, normalized to C 1s
area (green dots) that was set to 1.0. Solid lines represent tting results. Please note
that data points for SiOx and SiO2 at 20 are superimposed because they have the same
value. (For interpretation of the references to color in this gure legend, the reader is
referred to the web version of this article.)

Once obtained the CNT absorption coefcient on the basis of the


calibration provided by the AR-XPS analysis, we can easily estimate
the thickness of every CNT lms we realize without performing ARXPS measurements on each sample anymore, provided we measure
lm absorption spectra. Finally, we can tailor in a controlled fashion
the thickness of our CNT lms, simply by varying the aliquots of the
dispersion volume V m/c, where m is the CNT dispersed mass and
c the dispersion concentration. As plotted in Fig. 6, we found the
empirical law t C1 C2V, where C1, C2 are two t constants
dependent on the dispersion CNT and surfactant concentrations [1].
This is conrmed in Fig. 6, where the slopes of the two ts are
different, because we used two different dispersion concentrations
cMWCNT 0.003% w/v and cSWCNT 0.008% w/v for MWCNTs and
SWCNTs, respectively. We remark that, measuring the CNT lm
thickness by optical spectroscopy is a reliable and useful way to
estimate the effective lm thickness, neglecting its complex
microscopic fractal morphology [1].

32

F. De Nicola et al. / Carbon 95 (2015) 28e33

Fig. 4. (a) Absorption spectra of SWCNT and MWCNT lms for different aliquot volumes. (b) Estimation of SWCNT (blue dots) and MWCNT (red squares) absorption coefcients
through the LamberteBeer law. (c) Absorption coefcients of SWCNT (blue solid line) and MWCNT (red solid line) lms. Colored regions represent the rst S11, the second S22, and
the third S33 optical transition ranges for semiconducting SWCNT chirality, and the rst M11 range for metallic SWCNT chirality. The MWCNT lm has no signicant absorption peaks
but the p-p* surface plasmon polariton at z253 nm (5.1 eV), while for the SWCNT lm it is centered at z243 nm (4.9 eV). (d) Absorption lengths of SWCNT (blue solid line) and
MWCNT (red solid line). (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)

Fig. 5. Experimental absorption coefcient of conventional semiconductors taken


from Ref. [52] as a comparison with our SWCNT and MWCNT random network lms.

4. Conclusion
In summary, we fabricated SWCNT and MWCNT thin lms by
vacuum-ltration of aqueous dispersions. Such lms were deposited on substrates by dry-transfer printing at room temperature.
Furthermore, SEM images revealed the complex nature of CNT
random networks, which makes the estimation of their lm
thickness very difcult. Therefore, in order to obtain the CNT lm
thickness we performed AR-XPS measurements. Then, we discussed the optical spectra of CNT lms, once estimating their absorption coefcients with the LamberteBeer law and the obtained
AR-XPS lm thickness. Also, we provided an estimation of the CNT

Fig. 6. Thickness of SWCNT (blue dots) and MWCNT (red squares) lms as a function of
the dispersion aliquot volume. (For interpretation of the references to color in this
gure legend, the reader is referred to the web version of this article.)

lm absorption lengths. The latter have highlighted the importance


of CNT lms in light sensing and emitting applications. Finally, we
believe that our work may provide not only an useful, rapid, and
effective method to estimate the thickness of generic SWCNT and
MWCNT lms, but also a powerful tool to control the effective lm
thickness on the basis of the CNT dispersion volume.
Acknowledgments
We thank R. De Angelis, F. De Matteis, and P. Prosposito (Uni di Roma Tor Vergata, Roma, Italy) for their courtesy of
versita
spectrophotometry instrumentation. This project was nancial

F. De Nicola et al. / Carbon 95 (2015) 28e33

supported by the European Ofce of Aerospace Research and


Development (EOARD) through the Air Force Ofce of Scientic
Research Material Command, USAF, under Grant No. FA9550-14-10047.
References
[1] F. De Nicola, P. Castrucci, M. Scarselli, F. Nanni, I. Cacciotti, M. De Crescenzi,
Multi-fractal hierarchy of single-walled carbon nanotube hydrophobic coatings, Sci. Rep. 5 (2015) 1e9.
[2] F. De Nicola, P. Castrucci, M. Scarselli, F. Nanni, I. Cacciotti, M. De Crescenzi,
Exploiting the hierarchical morphology of single-walled and multi-walled
carbon nanotube lms for highly hydrophobic coatings, Beilstein J. Nanotechnol. 6 (2015) 353e360.
[3] P. Avouris, M. Freitag, V. Perebeinos, Carbon-nanotube photonics and optoelectronics, Nat. Photonics 2 (2008) 341e350.
ppler, C. Mann, T.C. Hain, F.M. Neubauer, G. Privitera, F. Bonaccorso,
[4] F. Scho
D. Chu, A.C. Ferrari, T. Hertel, Molar extinction coefcient of single-wall carbon
nanotubes, J. Phys. Chem. C 115 (2011) 14682e14686.
[5] M.F. Islam, D.E. Milkie, C.L. Kane, A.G. Yodh, J.M. Kikkawa, Direct measurement
of the polarized optical absorption cross section of single-wall carbon nanotubes, Phys. Rev. B 93 (2004) 0374041e0374044.
[6] H. Ye, X.J. Wang, W. Lin, C.P. Wong, Z.M. Zhang, Infrared absorption coefcients of vertically aligned carbon nanotube lms, Appl. Phys. Lett. 101
(2012) 1419091e1419094.
[7] G. Salvinelli, G. Drera, C. Baratto, A. Braga, L. Sangaletti, Stoichiometry
gradient, cation interdiffusion, and band alignment between a nanosized Tio2
blocking layer and a transparent conductive oxide in dye-sensitized solar cell
front contacts, Appl. Mater. Interfaces 7 (2015) 765e773.
[8] C. Pintossi, G. Salvinelli, G. Drera, S. Pagliara, L. Sangaletti, S. Del Gobbo,
M. Morbidoni, M. Scarselli, M. De Crescenzi, P. Castrucci, Direct evidence of
chemically inhomogeneous, nanostructured, Si-O buried interfaces and their
effect on the efciency of carbon nanotube/si photovoltaic heterojunctions,
J. Phys. Chem. C 117 (2013) 18688e18696.
[9] S.A. Chambers, M.H. Engelhard, V. Shutthanandan, Z. Zhu, T.C. Droubay,
P.V.S.L. Qiao, T. Feng, H.D. Lee, T. Gustafsson, E. Garfunkel, A. Shah, J.-M. Zuo,
Q.M. Ramasse, Instability, intermixing and electronic structure at the epitaxial
LaAlO3/SrTiO3(001) heterojunction, Surf. Sci. Rep. 65 (2010) 317e352.
[10] Y. Jia, J. Wei, K. Wang, A. Cao, Q. Shu, X. Gui, Y. Zhu, D. Zhuang, G. Zhang, B. Ma,
L. Wang, W. Liu, Z. Wang, J. Luo, D. Wu, Nanotube-silicon heterojunction solar
cells, Adv. Mater. 20 (2008) 4594e4598.
[11] M. Mohammed, Z. Li, J. Cui, T.-P. Chen, Acid-doped multi-wall carbon
nanotube/n-Si heterojunctions for enhanced light harvesting, Sol. Energy 106
(2014) 171e176.
[12] E. Shi, L. Zhang, Z. Li, P. Li, Y. Shang, Y. Jia, J. Wei, K. Wang, H. Zhu, D. Wu,
S. Zhang, A. Cao, TiO2-coated carbon nanotube-silicon solar cells with efciency of 15%, Sci. Rep. 2 (2012) 1e5.
[13] D.D. Tune, B.S. Flavel, R. Krupke, J.G. Shapter, Carbon nanotube-silicon solar
cells, Adv. Energy Mater. 2 (2012) 1043e1055.
[14] P. Castrucci, Carbon nanotube/silicon hybrid heterojunctions for photovoltaic
devices, Adv. Nano Res. 2 (2014) 23e56.
[15] S. Del Gobbo, P. Castrucci, S. Fedele, L. Riele, A. Convertino, M. Morbidoni, F. De
Nicola, M. Scarselli, L. Camilli, M. De Crescenzi, Silicon spectral response
extension through single wall carbon nanotubes in hybrid solar cells, J. Mater.
Chem. C 1 (2013) 6752e6758.
[16] Y. Jia, A. Cao, X. Bai, Z. Li, L. Zhang, N. Guo, J. Wei, K. Wang, H. Zhu, D. Wu,
P.M. Ajayan, Achieving high efciency silicon-carbon nanotube heterojunction solar cells by acid doping, Nano Lett. 11 (2011) 1901e1905.
[17] S.N. Habisreutinger, T. Leijtens, G.E. Eperon, S.D. Stranks, R.J. Nicholas,
H.J. Snaith, Carbon nanotube/polymer composites as a highly stable hole
collection layer in perovskite solar cells, Nano Lett. 14 (2014) 5561e5568.
[18] M.E. Itkis, F. Borondics, A. Yu, R.C. Haddon, Bolometric infrared photoresponse
of suspended single-walled carbon nanotube lms, Science 312 (2006)
413e416.
[19] J.H. Lehman, C. Engtrakul, T. Gennett, A.C. Dillon, Single-wall carbon nanotube
coating on a pyroelectric detector, Appl. Opt. 44 (2005) 483e488.
[20] Z.-P. Yang, L. Ci, J.A. Bur, S.-Y. Lin, P.M. Ajayan, Experimental observation of an
extremely dark material made by a low-density nanotube array, Nano Lett. 8
(2008) 446e451.
[21] E. Theocharous, C. Engtrakul, A.C. Dillon, J. Lehman, Infrared responsivity of a
pyroelectric detector with a single-wall carbon nanotube coating, Appl. Opt.
47 (2008) 3999e4003.
[22] K. Mizuno, J. Ishii, H. Kishida, Y. Hayamizu, S. Yasuda, D.N. Futaba, M. Yumura,
K. Hata, A black body absorber from vertically aligned single-walled carbon
nanotubes, PNAS 106 (2009) 6044e6047.
[23] J. Lehman, A. Sanders, L. Hanssen, B. Wilthan, J. Zeng, C. Jensen, Very black
infrared detector from vertically aligned carbon nanotubes and electric-eld
poling of lithium tantalate, Nano Lett. 10 (2010) 3261e3266.
[24] Z.-P. Yang, M.-L. Hsieh, J.A. Bur, L. Ci, L.M. Hanssen, B. Wilthan, P.M. Ajayan, S.Y. Lin, Experimental observation of extremely weak optical scattering from an
interlocking carbon nanotube array, Appl. Opt. 50 (2011) 1850e1855.

33

[25] H. Ye, X.J. Wang, W. Lin, C.P. Wong, Z.M. Zhang, Infrared absorption coefcients of vertically aligned carbon nanotube lms, Appl. Phys. Lett. 101
(2012) 1419091e1419094.
[26] A.B. Kaul, J.B. Coles, M. Eastwood, R.O. Green, P.R. Bandaru, Ultra-high optical
absorption efciency from the ultraviolet to the infrared using multi-walled
carbon nanotube ensembles, Small 9 (2013) 1058e1065.
[27] J.-Y. Kim, M. Kim, H. Kim, J. Joo, J.-H. Choi, Electrical and optical studies of
organic light emitting devices using SWCNTs-polymer nanocomposites, Opt.
Mater. 21 (2002) 147e151.
[28] K.W. Lee, S.P. Lee, H. Choi, K.H. Mo, J.W. Jang, H. Kweon, C.E. Lee, Enhanced
electroluminescence in polymer-nanotube composites, Appl. Phys. Lett. 91
(2007) 0231101e0231103.
[29] P. Fournet, J.N. Coleman, B. Lahr, A. Drury, W.J. Blau, D.F. O'Brien, H. rhold, Enhanced brightness in organic light-emitting diodes using a
H. Ho
carbon nanotube composite as an electron-transport layer, J. Appl. Phys. 90
(2001) 969e975.
[30] J.A. Misewich, R. Martel, P. Avouris, J.C. Tsang, S. Heinze, J. Tersoff, Electrically
induced optical emission from a carbon nanotube fet, Science 300 (2003)
783e786.
[31] M.P. Seah, I. Gilmore, G. Beamson, Xps: binding energy calibration of electron
spectrometers 5-re-evaluation of the reference energies, Surf. Interface Anal.
26 (1998) 642e649.
[32] J.J. Yeh, I. Lindau, Atomic subshell photoionization cross sections and asymmetry parameters: 1Z103, At. Data Nucl. Data Tables 32 (1985) 1e155.
[33] W.S.M. Werner, Electron transport in solids for quantitative surface analysis,
Surf. Int. Anal. 31 (2001) 141e176.
[34] A. Jablonski, Transport cross section for electrons at energies of surfacesensitive spectroscopies, Phys. Rev. B 58 (1998) 16470e16480.
[35] I.S. Tilinin, A. Jablonski, J. Zemek, S. Hucek, Escape probability of signal photoelectrons from non-crystalline solids: inuence of anisotropy of photoemission, J. Electron Spectrosc. Relat. Phenom. 97 (1997) 127e140.
[36] G. Drera, G. Salvinelli, J. hlund, P. Karlsson, B. Wannberg, E. Magnano,
L. Sangaletti, Transmission function calibration of an angular resolved
analyzer for X-ray photoemission spectroscopy: theory vs experiment,
J. Electron Spectrosc. Relat. Phenom. 195 (2013) 109e116.
[37] S. Tanuma, C.J. Powell, D.R. Penn, Calculation of electron inelastic mean free
paths (imfps) vii. reliability of the tpp-2m imfp predictive equation, Surf.
Interf. Anal. 35 (2003) 268e275.
[38] F.J. Himpsel, F.R. McFeely, A. Taleb-Ibrahimi, J.A. Yarmoff, G. Hollinger,
Microscopic structure of the SiO2/Si interface, Phys. Rev. B 38 (1988)
6084e6096.
[39] C. Pintossi, S. Pagliara, G. Drera, F. De Nicola, P. Castrucci, M. De Crescenzi,
M. Crivellari, M. Boscardin, L. Sangaletti, Steering the efciency of carbon
nanotube-silicon photovoltaic cells by acid vapor exposure: a real-time
spectroscopic tracking, ACS Appl. Mater. Interfaces 7 (2015) 9436e9444.
[40] A. Jorio, G. Dresselhaus, M.S. Dresselhaus, Carbon Nanotubes Advanced Topics
in the Synthesis, Structure, Properties and Applications, Springer, New York,
2008.
[41] L.A. Bursill, P.A. Stadelmann, J.I. Peng, S. Prawer, Surface plasmon observed for
carbon nanotubes, Phys. Rev. B 49 (1994) 2882e2887.
[42] P.M. Ajayan, S. Iijima, T. Ichihashi, Electron-energy-loss spectroscopy of carbon nanometer-size tubes, Phys. Rev. B 47 (1993) 6859e6862.
[43] B.W. Reed, M. Sarikaya, Electronic properties of carbon nanotubes by transmission electron energy-loss spectroscopy, Phys. Rev. B 64 (2001)
1954041e19540413.
[44] M.F. Lin, D.S. Chuu, C.S. Huang, Y.K. Lin, K.W.K. Shung, Collective excitations in
a single-layer carbon nanotube, Phys. Rev. B 53 (1996) 15493e15496.
[45] C. Kramberger, R. Hambach, C. Giorgetti, M.H. Rmmeli, M. Knupfer, J. Fink,
B. Bchner, L. Reining, E. Einarsson, S. Maruyama, F. Sottile, K. Hannewald,
V. Olevano, A.G. Marinopoulos, T. Pichler, Linear plasmon dispersion in singlewall carbon nanotubes and the collective excitation spectrum of graphene,
Phys. Rev. Lett. 100 (2008) 1968031e1968034.
[46] T. Pichler, M. Knupfer, M.S. Golden, J. Fink, A. Rinzler, R.E. Smalley, Localized
and delocalized electronic states in single-wall carbon nanotubes, Phys. Rev.
Lett. 80 (1998) 4729e4732.
[47] P. Castrucci, M. Scarselli, M.D. Crescenzi, M.A. El Khakani, F. Rosei, Probing the
electronic structure of carbon nanotubes by nanoscale spectroscopy, Nanoscale 2 (2010) 1611e1625.
[48] Y. Murakami, E. Einarsson, T. Edamura, S. Maruyama, Polarization dependence
of the optical absorption of single-walled carbon nanotubes, Phys. Rev. Lett.
94 (2005) 0874021e0874024.
[49] M. Lotya, Y. Hernandez, P.J. King, R.J. Smith, V. Nicolosi, L.S. Karlsson,
F.M. Blighe, S. De, Z. Wang, I.T. McGovern, G.S. Duesberg, J.N. Coleman, Liquid
phase production of graphene by exfoliation of graphite in surfactant/water
solutions, J. Am. Chem. Soc. 131 (2009) 3611e3620.
[50] A.B. Djurisi^
c, E.H. Li, Optical properties of graphite, J. Appl. Phys. 85 (1999)
7404e7410.
^telain, T. Gern, R. Humphrey-Baker, L. Forro,
[51] W.A. de Heer, W.S. Bacsa, A. Cha
D. Ugarte, Aligned carbon nanotube lms: production and optical and electronic properties, Science 268 (1995) 845e846.
[52] J. Nelson, The Physics of Solar Cells, Imperial College Press, London, 2003.

Vous aimerez peut-être aussi