You are on page 1of 6

2898

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

Creation of Individual Defects at Extremely High


Proton Fluences in Carbon Nanotube p n Diodes
Everett S. Comfort, Matthew Fishman, Argyrios Malapanis, Harold Hughes, Patrick McMarr, Cory D. Cress,
Hassaram Bakhru, and Ji Ung Lee

AbstractWe show that carbon nanotubes are robust under high

H+2 ion fluences. We draw this conclusion by analyzing radiation-

induced defects in reconfigurable single-walled carbon nanotube


p-n diodes with partially suspended nanotubes. Our analysis show
that any defects created through radiation is likely the result of
interactions between the nanotube and the substrate, whereas the
suspended region of the nanotube remains undamaged. In addition, we show that key features in the diode characteristics can be
explained by a single radiation-induced defect that enhances the
minority carrier generation rate of only one carrier type.
Index TermsCarbon nanotubes, defect analysis, p-n junctions,
proton irradiation.

I. INTRODUCTION

INCE their discovery, single-walled carbon nanotubes


(SWNTs) have attracted considerable attention for use in
future high-speed nanoelectronic devices. Unique properties
of SWNTs that are advantageous for nanoelectronics are very
high mobilities [1], and a bandgap that is tunable with both
diameter and chirality [2]. SWNTs have also recently attracted
interest for photovoltaics based on evidence of multiple exciton
generation from a single photon, seen in SWNTs configured as
diodes [3]. Recent progress in creating high percentage
semiconducting SWNT solutions through different sorting
techniques [4], [5] increases the likelihood of future commercial applications, which have significant potential for uses in
high-radiation environments. Because irradiating semiconducting materials can cause detrimental effects, such as lattice
defects that can introduce band gap states, research on the
robustness of SWNTs exposed to radiation is important. The
effects of proton [6], [7], electron [8] and ion radiation [9],
[10] on the morphology of carbon nanotubes have been studied
previously. Additionally, ionizing effects from these types of
radiation, as well as gamma radiation, have been studied using
Manuscript received July 22, 2011; revised September 13, 2011, September
26, 2011; accepted September 26, 2011. Date of publication October 26, 2011;
date of current version December 14, 2011. This work was supported in part
by the National Science Foundation under Grant EPDT-0823715, the Defense
Threat Reduction Agency under Grant HDTRA1-10-1-0016, and the NRL
under Grant N00173-10-1-G019.
E. Comfort, A. Malapanis, H. Bakhru, and J. U. Lee are with the College of
Nanoscale Science and Engineering, University of Albany-SUNY, Albany, NY
12203 USA (e-mail: jlee1@uamail.albany.edu).
M. Fishman is with Cornell University, Ithaca, NY 14840 USA.
H. Hughes, P. McMarr, and C. D. Cress are with the U.S. Naval Research
Laboratory, Washington, DC 20375 USA.
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TNS.2011.2170708

SWNTs configured as field effect transistors (FETs) [11], [12].


Nonvolatile memory based on carbon nanotubes have also been
studied, demonstrating the potential for radiation hardness in
carbon nanotube based nanoelectronics [13]. However, the
effects of radiation induced lattice damage, referred to as displacement damage, on SWNT electronic properties is difficult
to study using a FET configuration since the effects of ionizing
radiation may mask changes caused by displacement damage.
diodes to study the
Here, we use reconfigurable SWNT
ion irradiation on individual SWNTs
fundamental effects of
towards a greater understanding of radiation effects in SWNT
devices. This device structure allows us to probe the generation
of defects unambiguously. We show that SWNTs are very
robust to proton irradiation, with very few defects forming
along a single nanotube, even at the highest fluence of
protons/cm . We further examine the enhanced generation of
minority carriers caused by a single radiation-induced defect
and show that the defect enhances the minority carrier generation rate of one carrier type.
diode is a minority carrier based device, and therefore
A
diode,
very sensitive to any changes in their density. For a
relation is given by
the current-voltage
(1)
where is the charge on an electron,
is the voltage across
the source and drain, is the Boltzmann constant, is the absolute temperature, and is a measure of the ideality of the diode
(ideal diodes show
). is determined mainly by the density of minority carriers in both of the doped regions. Therefore,
any defect state created through irradiation could enhance the
minority carrier concentration above the background level. Ap, where
is the electron
proximately,
is the hole concentraconcentration in the -doped region,
is the diffusion length, and is
tion in the -doped region,
the carrier lifetime, assumed to be the same for both holes and
electrons. The change in minority carrier concentration caused
by radiation damage can be detected as an increase in , but we
diode provides additional
will show that a reconfigurable
information on defect density and the type of defect being generated not possible with diodes with fixed dopants, particularly
when analyzing materials with a low damage cross section.
II. EXPERIMENTS
A schematic of our SWNT
diode is shown in Fig. 1(a).
Its fabrication process has been reported elsewhere [14], and
is summarized below. First, 100 nm of SiO is deposited on a
silicon substrate. Next, gates used to electrostatically dope the

0018-9499/$26.00 2011 IEEE

COMFORT et al.: CREATION OF INDIVIDUAL DEFECTS AT EXTREMELY HIGH PROTON FLUENCES

Fig. 1. (a) Schematic of a p n diode device under proton irradiation. The total
SWNT length is between 6 and 10 m and the junction region varies from 0.3
to 2 m. (b) I V curves of as grown (open circles) and post-irradiation with a
fluence of 10 protons/cm (filled circles) are shown. A large increase for the
p n configuration after irradiation can be seen. The fits to the diode equation
are shown as solid lines. Before irradiation, the fitting parameters are I = 0:35
pA, n = 1:12, and R = 2:8 and 4.7 M
for the n p and p n configurations,
respectively. After irradiation, I = 0:63 pA, n = 1:2 and R = 10 M
for
the n p configuration, while the p n configuration can no longer be fit to the
diode equation.

carbon nanotube are created using 100 nm of deposited polysilicon, which is made highly conductive through phosphorous implantation and an activation anneal. The polysilicon gates are
defined using photolithography and reactive ion etching (RIE).
To electrically isolate the gates from the SWNT, SiO is deposited and planarized using chemical mechanical planarization to a final thickness of approximately 20 nm above the split
gates. Bondpads to the polysilicon are etched into the SiO , and
a 50 nm thick film of TiN is sputtered onto the wafer and patterned. TiN is chosen as it is compatible with the high temperature growth process of SWNTs. To suspend the SWNT, a trench
between the two split gates is defined and etched into the SiO
to a depth of
nm. PMMA is then spun onto the wafer
and patterned to define the SWNT catalyst. Finally, the PMMA
is lifted off in acetone, and SWNTs are grown using a catalytic
chemical vapor deposition process that is known to grow a wide
range of SWNT chiralities [15]. As our devices will encompass
many different chirality SWNTs, we note here that there is unlikely to be a chirality dependence to our results.
radiation source was used in our experiments.
A 2 MeV
Because an
ion dissociates into two protons upon impinging

2899

ion energy, the


on a target, each with half of the incident
radiation source used is expected to create effects similar to
a 1 MeV proton source. As such, we will refer to the equivalent proton dose throughout our analysis. The energy of the
ions is sufficiently high that the majority of the disassociated protons will be deposited well below our device structure, more than 10 m into the silicon substrate, determined
through SRIM simulations [16]. Over a dozen devices were irrato
diated, with the equivalent proton fluences ranging from
protons/cm . The highest fluence corresponds to approxiions impinging on a single nanotube. SWNT
mately
diode devices were electrically characterized using either a
Keithley 4200-SCS or Agilent B1500A parameter analyzer. We
behavior of a device biased as a
diode to (1)
fit the
and a series resistance
that takes into acand extract
count the nanotube-metal contact not captured in (1). Since the
SWNT is doped electrostatically, by switching the bias V and
or an
diode polarity
V on the split gates, either a
can be achieved with respect to fixed drain and source contacts.
configuration (V
V), a positive
In the
will forward bias the diode, while the same bias will reverse bias
configuration (V
V).
the diode in the
We also use two optical characterization methods to further
diodes.
understand the radiation response of SWNT
Raman spectra are taken before and after irradiation using 532
nm laser excitation. Raman spectroscopy is used to probe for
the characteristic D-band which is a measure of disorder in
SWNTs. In addition, photocurrent spectra, which provides a
unique signature of SWNT chirality, are also measured [17].
Using light from a quartz tungsten halogen lamp dispersed
through a monochromator, the photocurrent spectra of the
SWNT diodes are taken for photon energies from 0.52 eV.
This allows us to probe the excitonic transitions of a single
SWNT and band gap states that might be generated during irradiation. Electrical characterization, photocurrent spectroscopy
and Raman spectroscopy are all performed at room temperature
(300 K) in air.
III. RESULTS AND DISCUSSION
We irradiated many diodes for this study under a wide range
of fluences. Those that show defects show similar behavior. We
thus present detailed results from one diode as representative of
those that showed defects. We note that the majority of diodes
did not show evidence of defects, suggesting that the damage
cross-section of SWNTs is very small.
Before irradiation, our diode characteristics are nearly ideal
and are essentially identical for both configurations. In Fig. 1(b)
curves for one such diode for both
we show the diode
diode configurations (we discuss the effects due to irradiation
curve for each diode polarity is fit to (1) with
below). The
pA,
, and
a series resistance, resulting in
and 4.7 M for the
and
configurations,
respectively. We have shown previously that all diodes with
follow a universal behavior relating
to the lowest
exciton energy, suggesting that we are indeed probing the intrinsic properties of SWNTs [14].
Following initial characterization, the devices were irradiated
ions of varying fluences and again characterized
with 2 MeV

2900

Fig. 2. Ratio of I after irradiation to I before irradiation (I


=I )
is plotted vs. the proton fluence. Each bar represents one diode configuration.
The average ratio is also plotted versus fluence. A critical fluence of 10 protons/cm is needed to observe a change in diode behavior.

electrically. Devices exposed to a low fluence showed no sigbehavior and form the baseline
nificant degradation in the
for comparing to diodes exposed to a higher fluence, as summarized in Fig. 2. Only devices irradiated with fluences of
protons/cm or greater were observed to have an increase in ,
although not all were affected equally. Fig. 1(b) also shows the
curves after irradiation at a fluence of
protons/cm ,
configuration,
the highest fluence used in our study. In the
increases to 10 pA, more than ten times the initial value, and
the diode characteristics can no longer be reasonably fit to (1) for
configuration, on the other
any value of . The for the
hand, represents only a small change compared to the unirradiated case. The asymmetry shown in Fig. 1(b) is characteristic
of devices that showed degradation following irradiation, and is
used to carry-out our analysis for the type of defects being generated, discussed below. We note that the asymmetry between the
two diode configurations is in contrast to what we often observe
as SWNT diodes degrade over time with exposure to ambient
conditions, which always results in a symmetric increase in
for both configurations, presumably due to a relatively uniform
coverage of adsorbates onto the SWNT surface [18].
for the irradiated diodes
To explain the asymmetry in
we assert that a small number of defects are formed along
and values of 2 or
the SWNT, which can lead to increased
higher for [19][21]. Additionally, the inability of the
behavior to be fit to the ideal diode equation reflects a difference
in the recombination and generation processes following the
introduction of a defect. In this work, we choose to focus on
the effects of the defects on the generation process to describe
the increase in reverse bias current. Moreover, we later show
that the defects enhance the generation of only one minority
carrier type, which can be inferred from the asymmetry seen in
characteristics.
the
These conclusions are further supported by our results shown
in Fig. 2, which summarizes all of the devices and fluences

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

tested. Here we plot, for each fluence, the ratio of following


, and each
irradiation to the before irradiation
bar represents one diode configuration. For example, for a fluprotons/cm , three devices were tested, and both
ence of
and
configurations are plotted, resulting in the six
the
data points. Because some diodes can no longer be fit to (1) after
configuration shown in Fig. 1(b),
irradiation, such as the
the values are taken at a reverse bias of 200 mV to generate
the data in Fig. 2. In Fig. 2, we readily observe that a critical fluprotons/cm is needed to observe an increase in .
ence of
More importantly, in examining the higher fluence data, we obfor individual diodes is simserve that the ratio of
ilar in value for two decades increase in fluence. This suggests
that the number of defects along individual nanotubes, which
, does not scale with
should be proportional to
dose. We interpret this to mean that higher fluences only lead
to an increase in the number of diodes with increased , re.
sulting in the increase in the average values of
We note that only a single device was tested at the fluence of
protons/cm , but nonetheless shows the expected behavior
protons/cm . But
for devices above the critical fluence of
for all others, with the low damage cross section that seems to
be characteristic of our diodes, an increase in fluence increases
the probability of creating defects in our devices, but does not
significantly affect the induced defect density. Our conclusion is
consistent with some higher fluence diode configurations having
similar values to the lower fluence baseline. We are thus lead
to conclude that the rate of defect creation is very low, with a
lower limit possibly being one defect generated in diodes that
have been affected by irradiation. This is a good approximation
and consistent with the stochastic nature of defect creation in
the very low limit. This is further corroborated by noting that
a uniformly distributed defect density would necessarily lead
for both polarities, which is observed for
to an increase in
protons/cm ).
only one diode (the last data set irradiated at
Thus, the characteristics of this diode could be accounted for
by having defects in both doped sections of the nanotube. Our
analysis of the minority carrier generate rate, shown in Fig. 4,
further corroborates this conclusion.
Our conclusions based on diodes I-V characteristics are further corroborated by optical characterizations. Fig. 3(a) shows
the Raman spectra taken after irradiation (the data before irradiation is identical to that shown in Fig. 3(a)) for the device shown
in Fig. 1(b). Raman spectra are taken on the suspended portion
of the device where the signal is strongest; it is weak or often
not observed on the supported regions [22], [23]. In SWNTs, a
cm , called the D-peak, is a measure
peak centered at
of disorder. We note that this peak is not present in the spectra
even after being irradiated at the highest fluence. Likewise, no
D-peak was observed in all the other devices tested indicating
the robustness of SWNTs to proton irradiation. This, with the
additional optical characterization we discuss below, leads us to
conclude that defects are created on the supported regions.
In addition to the Raman spectra, we also measure the photocurrent spectra of individual diodes. The optical response of
SWNTs is dominated by excitonic transitions, unlike bulk semiconductors, and cannot be understood from a single-particle
framework. Excitons in SWNTs have large binding energies

COMFORT et al.: CREATION OF INDIVIDUAL DEFECTS AT EXTREMELY HIGH PROTON FLUENCES

Fig. 3. Optical charactization of a SWNT diode exposed to 10 protons/cm .


(a) Raman spectrum from the suspended junction region of the SWNT. The
spectrum shows no D-peak, usually located at 1350 cm , associated with defects in carbon nanotubes. (b) Photocurrent spectra of a device before and after
10 protons/cm irradiation. The E and E energies are 0.65 eV and 1.22
eV, respectively. Following irradiation, no new features emerge, again pointing
to a lack of damage to this portion of the SWNT.

of
meV [24], [25] and show stronger oscillator strength
than band-to-band transitions. The lowest optical transition is
referred to as the E exciton, the second the E exciton, etc.
When the energy of the photon is in resonance with an exciton
transition, a strong peak in the photocurrent is seen. The photocurrent spectra for the device in Fig. 1(b), both before and
after irradiation, are shown in Fig. 3(b). The photocurrent peaks
at 0.65 eV and 1.22 eV correspond to E and E transitions,
respectively. The spectra are normalized to photon flux and the
response after irradiation is offset for clarity. We can readily see
that no new features emerge after irradiation. Similar to Raman
spectroscopy, photoresponse is strongest in the region between
the split gates [3]. Because of the large binding energies of excitons in SWNTs, the spectra between exciton resonances probe
a significant portion of the band gap. Thus, the absence of any
new features after irradiation is another indication that the suspended portion of the SWNT remains undamaged.

2901

Our optical studies showing the absence of both new features


in the photocurrent spectra and a D-peak in the Raman spectra
indicate that no defects are created within the suspended region,
protons/cm . Also, since this part of
even at the fluence of
the nanotube is suspended well above the substrate, it will not
be affected by charge trapping or defect creation within the substrate below. These observations are consistent with our calcucm ,
lated differential proton cross-section of
prowhich results in negligible defects at the fluence of
tons/cm . Instead, we believe that defects are being produced
as a result of recoils from the SiO film that serves as the gate
dielectric for the split gates [26]. This conclusion is consistent
behavior, which we
with the asymmetry seen in the diode
explain by the introduction of defects within supported regions.
Based on this conclusion, we provide a detailed examination on
the minority carrier generation rate below.
The enhanced minority carrier generation rate can be examined using the Shockley-Read-Hall (SRH) model developed for
understanding generation and recombination in semiconductors
[27]. As our devices follow ideal diode behavior, the mechanisms that give rise to the forward and reverse bias current, recombination and generation respectively, are similar in nature
junction diodes. This justifies our
to a bulk semiconductor
use of SRH theory for our SWNT electrostatically doped
junctions. In addition, while minority and majority carriers are
generated in pairs in this process, the minority carrier density is
increased by orders of magnitude above the equilibrium density,
whereas the majority carrier concentration remains essentially
unchanged. We therefore use SRH theory to focus on the effects of a defect level on the minority carrier generation rate.
This phenomenological theory, however, is not completely applicable to analyzing a small number of defects. We can nevertheless use this model to provide the right framework towards
developing a quantitative analysis. More specifically, we use
SRH theory to determine the possible location of defects created in our device by examining the placement of defects either
within the junction region or one of the doped regions of our
device.
A centrally located defect in the junction region of our device would result in the same solution to the SRH equation for
both configurations. Since the solution is configuration indepencurves for this case,
dent, one would expect symmetric
and we therefore rule out this location. The asymmetry in the
curves, however, can be explained by defects located in
only one of the doped regions, supported by SiO , since the
solution of the SRH equation will be different depending on
the doping configuration. According to SRH theory, the generation rate from a -doped region, for example, is of the form
, where and
are the nonequilibrium and
is
equilibrium minority carrier densities, respectively, and
the electron lifetime [28]. If the value of is less than , as
would be expected near the junction of a
diode under reverse bias, will represent the generation of minority carriers.
For our devices, the doping is determined by the electric field
from the buried gates, which fixes the junction width for mod. Therefore, along the entire
erate applied biases
doped section the doping will remain constant, and the above
solution will hold. An equivalent relation exists for the -doped

2902

IEEE TRANSACTIONS ON NUCLEAR SCIENCE, VOL. 58, NO. 6, DECEMBER 2011

Greens function for Poissons equation in 1-D. The differential


equation we need to solve is given as

(2)

Fig. 4. Band diagrams illustrating the generation of minority carriers from a


single defect level, located at x . The edge of the field region is located at
x
. (a) When the SWNT is doped p-type, the minority carrier concentration
and defect G x; x
generated electrons.
is the sum of band-to-band n
(b) When the region is doped n-type, the defect no longer aids in minority carrier generation and the minority carrier concentration is only from band-to-band
 can result if the defect is posigeneration p . The inequality 
 if the defect is negatively
tively charged. A similar argument holds for 
charged.

=0

( )

( )

( (

))

region with a lifetime . Thus, asymmetry in the diode behavior


in symmetrically doped diodes can exist only when
,
explaining the dependence on doping configuration. Defects in
each of the doped regions, however, will result in increasing
for both configurations since the minority carrier generation will
be affected for both doping configurations. Although it might be
possible in any one device to have all the defects concentrated
to one side of the device this is highly unlikely for the many devices we have tested. Thus, the most plausible explanation for
characteristics is for
the observed asymmetry in the diode
there to be a very low defect density, possibly even a single defect, located in only one of the supported doped regions.
Although we cannot determine which minority carrier generation rate is enhanced, we can describe qualitatively the re. Later, we provide a plausible argument for
sults when
case, illustrated
this outcome. We now examine the
in Fig. 4, for a defect located at , as the doping is switched
from -type to -type. Under -type doping, the defect serves
as an efficient generation site, where minority carriers are generated and diffuse to the junction region, explaining the increase
for one of the configurations in Fig. 1(b). When this rein
gion is switched to -type doping, however, the defect does not
serve as an efficient generation site for holes, explaining the
small of change in for the other diode configuration shown in
. In the
Fig. 1(b). A similar argument can be made for
following treatment, we assume only a single defect within one
of the doped regions, but we cannot rule out the possibility of
more than one defect being present within this region. It is unlikely that the defect density is increased considerably, because
then it would become much more likely to create defects within
both doped regions, which will again result in a symmetric increase in . As such, it is very likely that it is possibly a single
defect created in one of the doped regions which gives rise to
behavior observed following irradiation.
the change in
To examine the effects from a single band gap state in a doped
region, we model the defect as a point source of minority carriers. In steady state, the diffusion equation with a source reduces to a Poisson equation. Our point source is modeled as
a Dirac delta function. This approximation is reasonable given
that nearly point-like defects can be generated in SWNTs under
ion damages [10]. Thus, the problem simplifies to solving the

is the diffusion coefficient,


is both the
Here,
Greens function and the minority carrier concentration that
results from the point-source, is the carrier lifetime, is
measured with respect to the edge of the field region, located
, and
denotes the location of the point source, as
at
illustrated in Fig. 4.
Equation (2) is subject to the usual boundary conditions that
is zero at the edge of field region and at the contact,
. The solution to (2) with these boundary conlocated at
ditions can be found by expanding in eigenfunctions and is given
as

(3)
can be found by solving the diffusion current at the edge of
the field region, given as
(4)
Remarkably, the diffusion current from a point source in a 1-D
system is independent of and is weakly dependent on for
most values. The weak dependence on the location of the defect
is reflected in the linear carrier concentration profile we find for
away from . In calculating
, we can simplify the
. The reverse bias current for this
sum by assuming
. Equating this to the leakage curcase is given as
rent, we arrive at the minority carrier lifetime of
ns. We
constrained
note that for this value of and for values of
within the doped region of our device, the minority carrier density where the defect is located can be at least an order of magnitude larger than the background minority carrier density for the
device in Fig. 1(b), validating the use of (2) [29]. The weak sento defect location can be further illustrated using
sitivity of
the value of calculated for the device in Fig. 1(a). Using this
value of , we calculate the change in for other defect locations. We find the ratio of
and
, indicating the
weak dependence of on . This weak dependence of
on
also shows that the asymmetry in the diode I-V characteristics cannot be explained simply by an asymmetric placement of
defects on both sides of the doped regions, again confirming the
single defect origin for the observed behavior. Finally, in the
absence of a detailed microscopic theory on the origin of our
defects, we only provide a qualitative explanation for the differ.) for the two diode configurations.
ence in the lifetimes (
The near absence of enhanced generation from the point source
for one of the diode configurations suggests that the defect is
charged. When a doped region is of the same charge type as
that of the defect, the free carriers screen the Coulomb potential
of the defect, but the defect still remains active as a generation

COMFORT et al.: CREATION OF INDIVIDUAL DEFECTS AT EXTREMELY HIGH PROTON FLUENCES

center for minority carriers. When doping is reversed so that the


doped region now has the opposite charge as that of the defect,
the defect is neutralized by the free carriers and is far less efficient in generating minority carriers.

IV. CONCLUSION
We studied the fundamental response of SWNTs to radiadiodes exposed to 2 MeV
tion using reconfigurable SWNT
ion radiation over a wide range of fluences. We used current-voltage characteristics and Raman and photocurrent spectroscopy to show that SWNTs are robust up to fluences of
protons/cm . For SWNTs damaged by radiation, we conclude
that as few as a single defect can describe many of the properties
of our diodes. The defect enhances the minority carrier generation rate of one type, and explains the characteristic asymmetry
characteristics. Defects are likely the result of
in the diode
substrate interactions with the SWNT. Based on this conclusion,
we believe future device applications for SWNTs in high radiation environments will require further study of substrate-nanotube interactions.

REFERENCES
[1] T. Drkop, S. A. Getty, E. Cobas, and M. S. Fuhrer, Extraordinary
mobility in semiconducting carbon nanotubes, Nano Lett., vol. 4, pp.
3539, 2003.
[2] T. W. Odom, J.-L. Huang, P. Kim, and C. M. Lieber, Atomic structure
and electronic properties of single-walled carbon nanotubes, Nature,
vol. 391, pp. 6264, 1998.
[3] N. M. Gabor, Z. Zhong, K. Bosnick, J. Park, and P. L. McEuen, Extremely efficient multiple electron-hole pair generation in carbon nanotube photodiodes, Science, vol. 325, pp. 13671371, 2009.
[4] M. Zheng, A. Jagota, M. S. Strano, A. P. Santos, P. Barone, S. Grace
Chou, B. A. Diner, M. S. Dresselhaus, R. S. Mclean, G. B. Onoa, G. G.
Samsonidze, E. S. Semke, M. Usrey, and D. J. Walls, Structure-based
carbon nanotube sorting by sequence-dependent DNA assembly, Science, vol. 302, pp. 15451548, 2003.
[5] M. S. Arnold, A. A. Green, J. F. Hulvat, S. I. Stupp, and M. C. Hersam,
Sorting carbon nanotubes by electronic structure using density differentiation, Nature Nanotechnol., vol. 1, pp. 6065, 2006.
[6] B. Khare, M. Meyyappan, M. H. Moore, P. Wilhite, H. Imanaka, and
B. Chen, Proton irradiation of carbon nanotubes, Nano Lett., vol. 3,
pp. 643646, 2003.
[7] V. A. Basiuk, K. Kobayashi, T. Kaneko, Y. Negishi, E. V. Basiuk, and
J.-M. Saniger-Blesa, Irradiation of single-walled carbon nanotubes
with high-energy protons, Nano Lett., vol. 2, pp. 789791, 2002.
[8] B. W. Smith and D. E. Luzzi, Electron irradiation effects in single wall
carbon nanotubes, J. Appl. Phys., vol. 90, pp. 35093515, 2001.
[9] V. Skkalov, A. B. Kaiser, U. Dettlaff, K. Arstila, A. V. Krasheninnikov, J. Keinonen, and S. Roth, Electrical properties of C irradiated
single-walled carbon nanotube paper, Phys. Status Sol. (B), vol. 245,
pp. 22802283, 2008.

2903

[10] A. Tolvanen, G. Buchs, P. Ruffieux, P. Grning, O. Grning, and A.


V. Krasheninnikov, Modifying the electronic structure of semiconducting single-walled carbon nanotubes by Ar ion irradiation, Phys.
Rev. B, vol. 79, p. 125430, 2009.
[11] C. D. Cress, J. J. McMorrow, J. T. Robinson, A. L. Friedman, and B. J.
Landi, Radiation effects in single-walled carbon nanotube thin-filmtransistors, IEEE Trans. Nucl. Sci., vol. 57, no. 6, pt. 1, pp. 30403045,
Dec. 2010.
[12] W.-K. Hong et al., Radiation hardness of the electrical properties
of carbon nanotube network field effect transistors under high-energy
proton irradiation, Nanotechnol., vol. 17, p. 5675, 2006.
[13] M. N. Lovellette, A. B. Campbell, H. L. Hughes, R. K. Lawerence, J.
W. Ward, M. Meinhold, T. R. Bengtson, G. F. Carleton, B. M. Segal,
and T. Rueckes, Nanotube memories for space applications, in Proc.
IEEE Aerosp. Conf., 2004, vol. 4, pp. 23002305.
[14] A. Malapanis, D. A. Jones, E. Comfort, and J. U. Lee, Measuring
carbon nanotube band gaps through leakage current and excitonic transitions of nanotube diodes, Nano Lett., vol. 11, pp. 19461951, 2011.
[15] J. Kong, H. T. Soh, A. M. Cassell, C. F. Quate, and H. Dai, Synthesis of individual single-walled carbon nanotubes on patterned silicon wafers, Nature, vol. 395, pp. 878881, 1998.
[16] J. F. Ziegler, Nuclear instruments and methods in physics research
section B: Beam interactions with materials and atoms, in Proc. SRIM,
2004, vol. 219, pp. 10271036.
[17] J. U. Lee, P. J. Codella, and M. Pietrzykowski, Direct probe of excitonic and continuum transitions in the photocurrent spectroscopy of
individual carbon nanotube p-n diodes, Appl. Phys. Lett., vol. 90, p.
053103, 2007.
[18] A. Malapanis, E. Comfort, and J. U. Lee, Current-induced cleaning
of adsorbates from suspended single-walled carbon nanotube diodes,
Appl. Phys. Lett., vol. 98, pp. 2631083, 2011.
[19] O. Breitenstein, J. Bauer, A. Lotnyk, and J. M. Wagner, Defect induced non-ideal dark I-V characteristics of solar cells, Superlattices
and Microstruct., vol. 45, pp. 182189, 2009.
[20] A. Schenk and U. Krumbein, Coupled defect-level recombination:
Theory and application to anomalous diode characteristics, J. Appl.
Phys., vol. 78, pp. 31853192, 1995.
[21] M. A. Trauwaert, J. Vanhellemont, E. Simoen, C. Claeys, B. Johlander,
L. Adams, and P. Clauws, Study of electrically active lattice defects
in Cf-252 and proton irradiated silicon diodes, IEEE Trans. Nucl. Sci.,
vol. 39, no. 6, pp. 17471753, Dec. 1992.
[22] Y. Kobayashi, T. Yamashita, Y. Ueno, O. Niwa, Y. Homma, and T.
Ogino, Extremely intense Raman signals from single-walled carbon
nanotubes suspended between Si nanopillars, Chem. Phys. Lett., vol.
386, pp. 153157, 2004.
[23] H. Son, Y. Hori, S. G. Chou, D. Nezich, G. G. Samsonidze, G. Dresselhaus, M. S. Dresselhaus, and E. B. Barros, Environment effects on the
Raman spectra of individual single-wall carbon nanotubes: Suspended
and grown on polycrystalline silicon, Appl. Phys. Lett., vol. 85, pp.
47444746, 2004.
[24] F. Wang, G. Dukovic, L. E. Brus, and T. F. Heinz, The optical resonances in carbon nanotubes arise from excitons, Science, vol. 308, pp.
838841, 2005.
[25] V. Perebeinos, J. Tersoff, and P. Avouris, Scaling of excitons in carbon
nanotubes, Phys. Rev. Lett., vol. 92, p. 257402, 2004.
[26] A. V. Krasheninnikov, K. Nordlund, and J. Keinonen, Production of
defects in supported carbon nanotubes under ion irradiation, Phys.
Rev. B, vol. 65, p. 165423, 2002.
[27] W. Shockley and W. T. Read, Statistics of the recombinations of holes
and electrons, Phys. Rev., vol. 87, p. 835, 1952.
[28] S. M. Sze, Physics of Semiconductor Devices. Hoboken, NJ: Wiley,
1981.
[29] J. U. Lee, Band-gap renormalization in carbon nanotubes: Origin of
the ideal diode behavior in carbon nanotube p-n structures, Phys. Rev.
B, vol. 75, p. 075409, 2007.

You might also like