Anda di halaman 1dari 17

Characterisation of the Growth Mechanism

during PECVD of Multiwalled Carbon


Nanotubes

Martin S. Bell1 , Rodrigo G. Lacerda2, Kenneth B.K. Teo1 , and


William I. Milne1
1
Engineering Department, University of Cambridge, Cambridge CB2 1PZ, UK
msb39@eng.cam.ac.uk
2
Universidade Federal de Minas Gerais, Departamento de Fı́sica, BR-30123-970
Belo Horizonte, MG, Brasil

Abstract. The growth mechanism of multiwalled carbon nanotubes has been a


subject of considerable research interest. Published results show that nanotubes
may be formed by a variety of methods using different combinations of carbon-
rich gases and etchant gases. We present an overview of different forms of carbon,
showing how nanotubes relate to other carbon structures and how they may be
produced. The use of plasma-enhanced chemical vapour deposition (PECVD) for
growing vertically aligned nanotubes for electronic device applications is reported,
including an analysis of the species present in the plasma during PECVD. It is
shown that the presence of a reactive species such as ammonia suppresses decom-
position of carbon feedstock gas and favours the formation of carbon nanotubes. It
is further shown that gas-phase reactions remove excess carbon, allowing the pro-
duction of nanotubes without unwanted amorphous carbon deposits, which is es-
sential for electronic applications. Plasma analysis is used to determine an optimum
carbon-source to etchant gas ratio, which is verified by postdeposition analysis.

1 Introduction
Carbon is a fascinating material due to the ability of carbon atoms to bond
with each other in different ways, giving rise to materials as varied as dia-
mond and graphite. Carbon has six electrons, arranged as (1s)2 (2s)2 (2p)2 .
It forms covalent bonds with other carbon atoms, using the outer four elec-
trons. To form these bonds, one of the 2s electrons is promoted to a 2p level.
The hybridisation of these electron orbitals allows carbon to form different
structures.
In diamond, sp3 hybridisation produces four identical and very strong
bonds. Each atom bonds to four others, giving rise to a tetrahedral structure,
symmetric in three dimensions.
In graphite, the remaining 2s electron hybridises with two of the 2p elec-
trons, giving three sp2 orbitals, which are located in an (x, y) plane, separated
by 120◦ . The remaining 2p electron orbital is perpendicular to this plane,
hence it is known as a pz orbital. The sp2 orbitals form strong σ-bonds be-

G. Messina, S. Santangelo (Eds.): Carbon, The Future Material for Advanced Technology Ap-
plications, Topics Appl. Phys. 100, 77–93 (2006)
© Springer-Verlag Berlin Heidelberg 2006
78 Martin S. Bell et al.

A
B
Fig. 1. AB stacking in graphite

Fig. 2. C60 molecule

tween the carbon atoms in the plane. Graphite consists of a series of these
parallel “graphene” planes.
The pz (or π) orbital perpendicular to the plane provides a weak van der
Waals bond that binds the planes together. The weakness of these π-bonds
allows the graphene planes to be easily separated. In high-quality “Bernal”
graphite, the graphene sheets are stacked ABAB. . . (Fig. 1).
These graphene sheets have an interlayer spacing of approximately
0.334 nm. The presence of defects in the graphene sheets gives rise to a larger
interlayer spacing (≈ 0.35 nm), and to a random rotation of the graphene
layers with respect to one another. This less-crystalline form is known as
turbostratic graphite.
A third structure observed in carbon is C60 , also known as “buckmin-
sterfullerene” after the architect Buckminster Fuller, who is famous for his
geodesic dome designs. The C60 molecule is an icosahedral structure made
up of 20 hexagons and 12 pentagons (Fig. 2). Each carbon atom is joined to
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 79

a2
a1

Fig. 3. Unit vectors in the graphene plane

three others, so it is predominantly σ-bonded, like graphite. There are similar


molecules containing more carbon atoms (e.g., C70 ).
The discovery of C60 in 1985 [1] may be identified as the first key step
along the path which led to the synthesis of carbon nanotubes. Carbon nano-
tubes themselves were identified in 1991 [2].
A single-walled nanotube is simply a graphene sheet rolled up into a
cylinder. The dangling bonds at the extremities of the sheet join together,
completing the cylinder. The size of the nanotube and its electrical properties
are determined by the “chiral vector”, which links the two points in the
graphene sheet that join up to complete the nanotube. The chiral vector is
defined as c = ma1 + na2 , where a1 and a2 are the unit vectors in the
graphene plane (Fig. 3).
The nanotube, of course, requires not just sides but a cap at the end
to eliminate the dangling bonds. The cap is an icosahedral structure made
up of pentagons and hexagons, like a C60 molecule. A nanotube is classified
according to its chiral vector, and referred to as an (m, n) tube. There are
two special cases that exhibit particularly high symmetry. These are known
as “armchair” and “zig-zag” nanotubes. For an armchair nanotube, n = m,
whilst for a zig-zag tube, n = 0. In these cases, the chiral vector is perpen-
dicular to the nanotube axis (Fig. 4). For the two specific nanotubes shown,
the cap in each case is half of a C60 molecule.
More generally, if n = 0, and n = m, a “chiral” nanotube is formed, where
the hexagons are arranged helically around the nanotube. Single-walled nano-
tubes typically have diameters in the range 1–5 nm. They may be metallic
(0 eV band gap) or semiconducting (typically 0.4–0.7 eV band gap) depend-
ing upon the chiral vector. The band gap is also inversely proportional to the
diameter [3, 4, 5].
A multiwalled nanotube consists, in essence, of a series of concentric single
walled nanotubes. As discussed above, the interlayer spacing in well-ordered
ABAB graphite is 0.334 nm. For the ABAB structure to be replicated in a
multiwalled nanotube would require successive nanotube circumferences to
80 Martin S. Bell et al.

Fig. 4. “Armchair (5,5)” and “zig-zag (9,0)” nanotubes (top and bottom, respec-
tively)

differ by an integer multiple of (2π × 0.334 nm), as well as requiring well-


ordered individual nanotubes.
Measurements have shown that the intershell spacing in multiwalled nano-
tubes can range from 0.34 to 0.39 nm, with the spacing decreasing as nano-
tube diameter increases [3, 6]. The geometrical constraints in forming the
seamless graphene cylinders cause the layers to be uncorrelated with respect
to one another. Multiwalled nanotubes therefore resemble the less-ordered
turbostratic form of graphite rather than the higher-quality Bernal form.
Multiwalled nanotubes typically have outer diameters in the range 50–
100 nm. Electrical conduction can take place within each of the individual
shells of a multiwalled nanotube. However, if a multiwalled nanotube is con-
tacted on the outside, the electric current is conducted through its outermost
shell only [7]. Where conduction occurs through this outer shell, the band
gap approaches 0 eV due to the relatively large diameter and the nanotube
acts as a metallic conductor.
Before we go any further, it is worth mentioning a related structure,
namely carbon nanofibres. Carbon nanofibres are another kind of graphitic fil-
ament, which differ from nanotubes in the orientation of the graphene planes.
The graphene planes in a carbon nanotube are parallel to the nanotube axis;
in a nanofibre, the graphene layers are not parallel to the axis, but are rather
arranged in a stacked form (layers perpendicular to the fibre axis) or herring-
bone form (layers at an angle to the axis). These nanofibres are illustrated
in Fig. 5, with the nanofibre axis indicated by an arrow in each case.
Nanofibres have many useful properties in common with carbon nano-
tubes and share some potential applications. They do not, however, share all
of the beneficial properties of nanotubes that derive from their structure.
When carbon nanotubes were discovered, it was realised that because
the graphene planes are parallel to the filament axis, they would inherit
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 81

Fig. 5. “Stacked” and “herringbone” nanofibres (top and bottom, respectively)

several important properties of graphite. In particular, a nanotube exhibits


high electrical conductivity, thermal conductivity and mechanical strength
along its axis. As there are very few open edges and dangling bonds in the
structure, nanotubes are also very inert, and species tend to be physically
adsorbed onto the graphene walls rather than react with them. As carbon
nanotubes are completely covalently bonded, they can be excellent electrical
conductors which do not suffer from electromigration, a limiting factor for
metals.
Carbon nanotubes have been investigated for a wide range of applications.
A nanotube can behave as a high-aspect-ratio electrical conductor with mi-
cron-scale length and nanometre-scale diameter. Structures of this type are
highly desirable as field emission tips for applications such as field emission
displays [8, 9], X-ray tubes [10], electron sources for microscopy and lithogra-
phy [11], gas discharge tubes [12] and vacuum microwave amplifiers. The high
aspect ratio and small diameter of the nanotube is also useful for scanning
probe tips [13, 14].
Semiconducting single-wall carbon nanotubes have been investigated for
use as transistors or logic elements [15, 16]. The electronic properties of the
carbon nanotubes are highly sensitive to adsorbed species [17, 18], which is a
challenge for logic circuits, requiring the nanotubes to be suitably encapsu-
lated. However, this sensitivity may be utilized in chemical or biological sen-
sors for gas detection. The coherent nature of electron transport in crystalline
nanotubes makes these structures potentially suitable for spin-electronic de-
vices [19].
Carbon nanotubes may also be used as electromechanical sensors as their
electrical characteristics change in response to mechanical deformation of
their structure [20]. Conversely, nanotubes can mechanically deflect under
electric stimulation (e.g., due to charge induced on the nanotubes) making
them useful as cantilevers or actuators [21, 22].
Aside from applications which exploit the properties of individual nano-
tubes, there are a further range of applications which utilise the large surface
82 Martin S. Bell et al.

area to volume ratio of carbon nanotubes and nanofibres when manufactured


in bulk. One interesting application for these structures is as electrodes in
electrochemical supercapacitors [23, 24]. It has been shown that such struc-
tures could lead to higher charge storage than conventional capacitors and
batteries [21, 25]. The high electrical conductivity and relative inertness of
nanotubes also make them potentially useful as electrodes in electrochemical
reactions [26, 27]. Nanotubes can support reactant particles in catalytic con-
version reactions [28, 29]. It has been proposed that hydrogen could be stored
amongst and inside nanotubes/nanofibres for fuel cell applications [29, 30],
although results show that the amount of hydrogen stored is not as high as
originally anticipated [31]. The use of nanotubes and nanofibres as filters or
membranes for molecular transport has also been proposed [32].
Nanotubes and nanofibres can be added to polymers to improve their
strength and stiffness, and can also add electrical conductivity to polymer-
based composite systems [33, 34], which has proved useful in the automobile
industry for the electrostatic painting of components. Carbon nanotubes are
ideal reinforcing fibres for composites due to their high aspect ratio and
axial strength [35]. Single-wall carbon nanotubes are preferred to multiwall
nanotubes because the inner layers of the multiwall nanotubes contribute
little under structural loading, and thus would reduce the stiffness for a given
volume of nanotubes [35, 36].

2 Production of Carbon Nanotubes


There are a number of known methods for producing carbon nanotubes.
Between them it is possible to produce nanotubes with differing properties
and in different forms. The most common methods used for the production
of nanotubes are arc discharge [37, 38], laser vaporisation [39] and chemical
vapour deposition.
The first nanotubes were produced in an arc discharge arrangement [2].
In a typical arc discharge apparatus, two graphite electrodes are held a short
distance apart inside a vacuum chamber. The chamber is held at high vac-
uum and a flow of helium is supplied. A DC voltage is applied across the two
electrodes, and a large arc current flows, removing carbon atoms from the
anode in the process. The position of the anode is adjustable so that as the
anode is consumed, the gap between the two electrodes can be held constant.
A carbon deposit forms on the cathode; this deposit contains nanotubes as
well as larger carbon nanoparticles. In a high-quality arc discharge, the de-
posit can contain 70% nanotubes. These nanotubes typically exhibit a good
crystalline structure and are very straight, due to the very high temperatures
(3000◦ C) developed during the arc discharge. However, in order to make use
of these nanotubes in an electronic device, they must be separated from the
cathode, purified and manipulated onto a substrate.
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 83

Like arc discharge, laser vaporisation is a short-duration, high-temper-


ature technique for nanotube production. A graphite target is heated to
around 1200◦C in a horizontal furnace whilst an inert gas flows across it.
A high-power laser is then used to vaporise the graphite target. The vapor-
ised material is carried by the gas onto a cooled collector, where it condenses.
As with arc discharge, the material contains a mixture of nanotubes and
nanoparticles, and once again the material must be separated, purified and
manipulated in order to fabricate devices. Laser vaporisation can produce
up to 90% nanotubes, however, unlike arc discharge, its use is limited to
small-scale production.
In both of these cases, a block of solid graphitic carbon is heated to
a very high temperature and carbon atoms are separated from the block,
reassembling either on the cathode in the case of arc discharge or on a cooled
collector in the case of laser vaporisation. It is during this reassembly that
the highly ordered nanotubes are formed.
Catalytic chemical vapour deposition is very different. Instead of begin-
ning with a block of carbon, the feedstock is a hydrocarbon gas, which dissoci-
ates either thermally (thermal CVD) or in the presence of a plasma (plasma-
enhanced CVD). Once again, the dissociated carbon atoms self-assemble into
highly ordered nanotubes; however, in this case the nanotubes form on a
substrate, which may be made of any material which can withstand the de-
position temperature (e.g., silicon or glass). The self-assembly is facilitated
by catalyst nanoparticles deposited on the substrate surface, which seed the
nanotube growth. These catalyst nanoparticles determine both the location
and diameter of the nanotubes formed. This ability to deposit nanotubes
selectively based upon the location of catalyst particles opens the door to
lithography and a range of possibilities for direct deposition of nanotube de-
vices.
Thermal CVD requires high temperatures (800–1000◦C) and has two dis-
advantages. First, the nanotubes produced are not just randomly oriented,
but also they are not straight. Second, the high temperature rules out the
use of many desirable substrate materials (e.g., glass).
Plasma-enhanced CVD (‘PECVD’) is a technique used extensively in the
semiconductor industry to allow CVD at lower temperatures. In this case, en-
ergy in the plasma replaces some of the heat energy, allowing gas dissociation
and nanotube formation to take place at lower temperatures (600–700◦C).
PECVD has a further advantage: the electric field aligns the nanotubes dur-
ing growth [40, 41]. There are a number of different techniques available for
creating the plasma. These include rf-PECVD [42], microwave PECVD [43],
inductively coupled PECVD [26] and DC glow discharge PECVD [44].
For an electronic device, it is essential to be able to make good electrical
contact to the nanotubes. Silicon is often selected as an appropriate substrate.
Also, for an electronic device, it is important that the synthesis of nanotubes
is performed without the deposition of amorphous carbon (“a-C”), which
84 Martin S. Bell et al.

prevents the formation of nanotubes by poisoning the growth catalyst and


can cause short circuits on the substrate surface. One method that has been
shown to prevent the deposition of a-C is to combine the carbon source (a
hydrocarbon gas) with a hydrogen-rich gas (typically NH3 or H2 ), which
produces reactive species in the plasma that remove any excess carbon.

3 Plasma Composition during PECVD


We have seen that we can form carbon nanotubes using several different
techniques. In the case of arc discharge and laser vaporisation, carbon is
liberated from a graphite source before reassembling in nanotube form. In
the case of CVD, it is not clear if carbon itself is present in the chemical
vapour, or if carbon is extracted from hydrocarbon molecules in a reaction
at the surface of the metal catalyst.
Different conditions for nanotube synthesis yield different results: Under
certain conditions we get well-defined nanotubes; at other conditions we get
amorphous carbon. What is the mechanism that determines this? In order to
answer these questions, we conducted an analysis of the plasma composition
during PECVD of carbon nanotubes [45].
Multiwalled carbon nanotubes were grown using a DC glow discharge
PECVD setup with the substrate located on a resistively heated graphite
stage. It has been reported that C2 H2 is the most efficient carbon feedstock
for the growth of carbon filaments [46], and we selected C2 H2 to be the carbon
feedstock gas for our carbon nanotube growth. NH3 was added to generate
reactive species, which have been shown to help prevent formation of a-C.
The gases were fed into the chamber through a metal pipe that acted as an
anode for the plasma discharge. A DC plasma was generated between this
pipe and the tungsten heaters within the graphite stage. The chamber was
maintained at vacuum using a rotary pump.
Si substrates coated with a thin catalyst film were placed on the graphite
stage and heated to 520◦ C in NH3 . At this temperature, the thin film cat-
alyst agglomerates into particles suitable for seeding nanotube growth. The
DC plasma was then initiated and C2 H2 added into the gas flow as the carbon
feedstock. The temperature was maintained at 650◦C throughout the deposi-
tion, and the chamber pressure was maintained at 5.3 mbar. The plasma DC
bias was maintained at 600 V, with current typically around 100 mA. The
volume flow-rate proportion of C2 H2 in the NH3 :C2 H2 plasma was varied
between 0% and 70% whilst maintaining other parameters at their standard
settings.
Mass spectrometry was performed on species extracted 15 mm from
the graphite stage. Neutral molecules were extracted from the plasma and
analysed using residual gas analysis (RGA). Positive ions were extracted and
analysed using secondary-ion mass spectrometry. An RGA mass spectrum
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 85

Fig. 6. RGA and positive ion mass spectra (top and bottom, respectively)

for neutral species and a mass spectrum for positive ion species at 23% C2 H2
are shown in Fig. 6.
The species in the RGA spectrum are consistent with the cracking pat-
terns of C2 H2 (base peak at amu 26) and NH3 (base peak at 17 amu) and also
indicate the presence of H2 , HCN, H2 O, N2 and CO2 [47]. This is consistent
with data reported by other authors [48]. The dominant ions detected were
NH+ + + + +
3 and C2 H2 . Other detected species were NH2 , NH4 , HCN and C2 H .
+

4 Characterisation of the Growth Mechanism


The results obtained by growing carbon nanotubes in this particular config-
uration have previously been reported [49]. It was demonstrated that well-
aligned nanotubes were grown for C2 H2 concentrations between 4–20%, that
at 29% C2 H2 the nanotubes became more obelisk-like, and that by 38% there
are no longer any tubes at all. These results are shown in Fig. 7. It was also
reported that the nanotube growth rate peaked at around 20% C2 H2 content.
This growth condition gave a clean, a-C-free substrate [50].
We first investigated the nature of the carbon precursor for the formation
of nanotubes. We were unable to detect C2 , CH4 or other higher carbon
species even after acquiring data for a prolonged period, and thus concluded
that C2 H2 is the dominant precursor for nanotube formation. This conclusion
is supported by other authors [51], who have detected the presence of C2 H2
86 Martin S. Bell et al.

Fig. 7. Scanning electron microscope images of carbon nanotubes grown at different


gas ratios

catalyst C xH y
particle C

C C

C C C xH y catalyst
Cx Hy
particle
barrier layer barrier layer

(a) weak interaction (b) strong interaction

Fig. 8. Growth models

in a CH4 /H2 /NH3 plasma yielding nanotubes. For this to be the case, our
growth model must include a reaction at the catalyst surface to extract carbon
from C2 H2 .
The catalyst takes the form of a transition metal (Fe, Ni, Co or Mo), which
may be applied chemically from a solution containing the catalyst [52] or
directly by using techniques such as thermal evaporation, ion beam sputter-
ing [53] or magnetron sputtering [54]. For our experiment we used magnetron
sputtering to give us a well-controlled Ni film.
When heat is applied to the substrate bearing the film, the increased sur-
face mobility of the catalyst atoms causes the film to coalesce into nanoclus-
ters [55]. The thickness of the catalyst film, temperature and time determine
the size of these nanoclusters [49, 56, 57].
There is a further consideration with respect to the catalyst layer. This
is the chemical interaction between the catalyst and substrate when heat is
applied. If there is a reaction, the catalyst material will dissipate, ending its
usefulness for seeding nanotube growth. This is a problem when using Si as
a substrate with transition metals, as they would diffuse into the substrate
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 87

at the temperatures used. In order to eliminate this problem, a thin diffusion


barrier of SiO2 was sputtered onto the Si substrate before the metal catalyst
layer was deposited.
Depending upon the strength of the interaction between the catalyst metal
and the barrier layer upon which it is deposited, two different growth models
have been proposed [58]. These two models are illustrated in Fig. 8. In each
case, carbon is extracted from the hydrocarbon feedstock gas and diffuses
through the catalyst particle before taking its place in the forming nanotube.
The growth model when there is weak interaction between the catalyst par-
ticle and the barrier layer is known as “tip growth”, as the catalyst particle
stays at the tip of the nanotube. The model when there is strong interaction
is known as “base growth”, as the catalyst particle remains anchored to the
barrier layer at the base of the nanotube. For nickel on SiO2 , interaction is
weak, indicating tip growth. Indeed, bright catalyst particles can clearly be
seen at the tips of the nanotubes in Fig. 7.
These growth models are widely accepted in the nanotube community,
but there is a puzzle concerning the diffusion of carbon through the catalyst
particle. The temperatures achieved during PECVD are well below the melt-
ing point of the catalyst metals, yet it is thought that the catalyst particle
must be in the liquid phase for this diffusion to take place. An explanation for
this is that the saturation of the metal with carbon atoms causes a significant
decrease in the melting point for the catalyst particles, allowing the particle
to melt and the diffusion to take place [59].
Figure 7 shows that varying the gas ratio produces different structures,
with carbon nanotubes deposited at low C2 H2 ratios and a-C deposited at
high C2 H2 ratios. In order to understand why this is the case, we must ex-
amine the role of NH3 in the plasma. The production of a-C-free nanotubes
requires a controlled deposition of carbon, which can then self-assemble into
an energetically favoured nanotube form. This controlled deposition rate is
achieved through the combination of two reactions: the dissociation of a
carbon-rich gas (in our case C2 H2 ) and the removal of excess carbon that
would otherwise lead to amorphous carbon deposits.
It has been widely reported that atomic hydrogen is the active species
for the removal of excess carbon. Both H2 and NH3 have been reported to
act as sources for atomic hydrogen through electron impact dissociation in
plasma [51, 60]. Generation of H from an NH3 :C2 H2 DC plasma has also
been shown in simulations [61]. It has further been shown that NH3 is a
more effective generator of atomic hydrogen than H2 [51]. NH3 therefore has
a key role in removing any excess carbon through the generation of reactive
species which combine with and carry away carbon atoms.
The efficacy of NH3 in nanotube production has been further attributed
to the preferential decomposition of NH3 suppressing the decomposition of
C2 H2 . This is because the chemical bonds that hold the NH3 molecule to-
gether are weaker than those which hold C2 H2 together. It has also been
88 Martin S. Bell et al.

Fig. 9. H2 RGA signal for varying gas ratio

Fig. 10. HCN RGA signal for varying gas ratio

suggested that nitrogen is incorporated into catalyst particles, forming ni-


trides and changing the composition of the catalyst surface [62]. This may
also help to explain the diffusion puzzle. We can clearly see the preferential
decomposition of NH3 in our RGA data. A graph of the RGA signal for H2
for varying gas ratio is shown in Fig. 9.
H2 is generated by the decomposition of both NH3 and C2 H2 . To the
left of the figure, where the plasma is predominantly NH3 , the amount of H2
increases as NH3 increases. In this region, the H2 is derived from decompo-
sition of NH3 . To the right of the figure, where the plasma is predominantly
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 89

C2 H2 , the amount of H2 increases as C2 H2 increases. In this region, the H2


is derived from the decomposition of C2 H2 .
At high NH3 ratios, NH3 decomposes preferentially over C2 H2 . This
causes the C2 H2 to decompose slowly, generating the small amounts of car-
bon necessary for nanotubes to self-assemble. At high C2 H2 ratios, there
is insufficient NH3 to effectively suppress C2 H2 decomposition, resulting in
high levels of carbon generation and deposition of excess carbon as a-C. NH3
therefore has two key roles in the formation of carbon nanotubes: not only
does it generate atomic hydrogen species to remove any excess carbon, but
it also suppresses the decomposition of C2 H2 , limiting the amount of carbon
generated in the first place.
Our observed minimum in H2 is close to our previously reported peak
in clean nanotube growth rate at around 20% C2 H2 . At this ratio, carbon-
removing species are at a minimum and C2 H2 decomposition is low, giving
rise to controlled deposition of clean nanotubes. Interestingly, ion intensities
peak at around 40% C2 H2 , where a-C is deposited. The optimum condition
for nanotube growth is therefore not the condition of maximum ion inten-
sity. This contrasts with most thin film growth, where the condition with
maximum ionisation is preferred.
Further work [63] has shown that the plasma during the nanotube depo-
sition is sufficiently energetic for endothermic ion-molecule reactions to take
place. In many applications, these reactions can and have been ignored as
they require significant pressures and ion energies. However, in this case the
high pressure and DC bias present the perfect condition for these reactions.
The process for the removal of excess carbon by atomic hydrogen species
should give rise to products including both carbon and hydrogen. The princi-
pal reaction product we detected in our RGA data was HCN. A graph of the
RGA signal for HCN with varying gas ratio is shown in Fig. 10. This shows
that HCN is generated where both C2 H2 and NH3 are in abundance.
The Figure also shows that a purely NH3 plasma does not generate signif-
icant HCN. This confirms that the reactions that remove excess carbon are
predominantly taking place in the gas phase. If there was a strong surface
reaction, we would see HCN arising from reactions between the active species
and the graphite stage, even in the absence of C2 H2 .

5 Conclusion
We have shown that carbon nanotubes have many exciting features and po-
tential applications. Nanotubes can be made using several different tech-
niques, and PECVD is particularly suitable for the fabrication of electronic
devices.
During PECVD, carbon is extracted from C2 H2 in a gas-phase reaction
at the catalyst surface, where it self-assembles into nanotube form. We have
90 Martin S. Bell et al.

explained why varying the gas ratio during deposition yields different struc-
tures and shown that there is an optimum ratio for clean nanotube growth.
This enables the fabrication of a-C-free devices for electronics applications,
which we believe opens up an enormous range of potential applications.
A vast amount of research effort has been conducted into carbon nano-
tubes, and we have an increasingly clear understanding of their properties
and how these can be manipulated to give desirable devices. Whilst nano-
tubes have made only limited forays into commercial products to date, we
believe that over the next decade the remarkable promise they have shown
in the research arena will be realised in the wider world.

References
[1] H. W. Kroto, J. R. Heath, S. C. O’Brien, R. F. Curl, R. E. Smalley: Nature
318, 162 (1985) 79
[2] S. Ijima: Nature (London) 354, 56 (1991) 79, 82
[3] M. Endo, K. Takeuchi, T. Hiraoka, T. Furuta, T. Kasai, X. Sun, C. H. Kiang,
M. S. Dresselhaus: J. Phys. Chem. Solids 58, 1707 (1997) 79, 80
[4] T. W. Odom, J. L. Huang, P. Kim, C. M. Lieber: Nature 391, 62 (1998) 79
[5] J. W. G. Wildoer, L. C. Venema, A. G. Rinzler, R. E. Smalley, C. Dekker:
Nature 391, 59 (1998) 79
[6] C. H. Kiang, M. Endo, P. M. Ajayan, G. Dresselhaus, M. S. Dresselhaus: Phys.
Rev. Lett. 81, 1869 (1998) 80
[7] A. Bachtold, C. Strunk, J. P. Salvetat, J. M. Bonard, L. Forro, T. Nussbaumer,
C. Schonenberger: Nature 397, 673 (1999) 80
[8] W. B. Choi, D. S. Chung, J. H. Kang, H. Y. Kim, Y. W. Jin, I. T. Han,
Y. H. Lee, J. E. Jung, N. S. Lee, G. S. Park, J. M. Kim: Appl. Phys. Lett. 75,
3129 (1999) 81
[9] D. S. Chung, S. H. Park, H. W. Lee: Appl. Phys. Lett. 80, 4045 (2002) 81
[10] G. Z. Yue, Q. Qiu, B. Gao, Y. Cheng, J. Zhang, H. Shimoda, S. Chang, J. P. Lu,
O. Zhou: Appl. Phys. Lett. 81, 355 (2002) 81
[11] N. deJonge, Y. Lanny, K. Schoots, T. H. Ooesterkamp: Nature 420, 393 (2002)
81
[12] R. Rosen, W. Simendinger, C. Debbault, H. Shimoda, L. Fleming, B. Stoner,
O. Zhou: Appl. Phys. Lett. 76, 1668 (2000) 81
[13] H. J. Dai, J. H. Hafner, A. G. Rinzler, D. T. Colbert, R. E. Smalley: Nature
384, 147 (1996) 81
[14] H. Nishijima, S. Kamo, S. Akita, Y. Nakayama, K. I. Hohmura,
S. H. Yoshimura, K. Takeyasu: Appl. Phys. Lett. 74, 4061 (1999) 81
[15] V. Derycke, R. Martel, J. Appenzeller, P. Avouris: Nano Lett. 1, 453 (2001)
81
[16] A. Bachtold, P. Hadley, T. Nakanishi, C. Dekker: Science 294, 1317 (2001) 81
[17] J. Kong, N. R. Franklin, C. W. Zhou, M. G. Chapline, S. Peng, K. J. Cho,
H. J. Dai: Science 287, 622 (2000) 81
[18] P. G. Collins, K. Bradley, M. Ishigami, A. Zettl: Science 287, 1801 (2000) 81
[19] K. Tsukagoshi, B. W. Alphenaar, H. Ago: Nature 401, 572 (1999) 81
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 91

[20] T. W. Tombler, C. W. Zhou, L. Alexseyev, J. Kong, H. J. Dai, L. Lei,


C. S. Jayanthi, M. J. Tang, S. Y. Wu: Nature 405, 769 (2000) 81
[21] R. H. Baughman, A. A. Zakhidov, W. A. de Heer: Science 297, 787 (2002)
81, 82
[22] R. H. Baughman, C. X. Cui, A. A. Zakhidov: Science 284, 1340 (1999) 81
[23] K. H. An, W. S. Kim, Y. S. Park, Y. C. Choi, S. M. Lee, D. C. Chung, D. J. Bae,
S. C. Lim, Y. H. Lee: Adv. Mater. 13, 497 (2001) 82
[24] M. Hughes, M. S. P. Shaffer, A. C. Renouf, C. Singh, G. Z. Chen, J. Fray,
A. H. Windle: Adv. Mater. 14, 382 (2002) 82
[25] E. Frackowiak, F. Beguin: Carbon 40, 1775 (2002) 82
[26] J. Li, R. Stevens, L. Delzeit, H. T. Ng, A. Cassell, J. Han, M. Meyyappan:
Appl. Phys. Lett. 81, 910 (2002) 82, 83
[27] M. A. Guillorn, T. E. McKnight, A. Melechko, V. I. Merkulov, P. F. Britt,
D. W. Austin, D. H. Lowndes, M. L. Simpson: J. Appl. Phys. 91, 3824 (2002)
82
[28] G. L. Che, B. B. Lakshmi, E. R. Fisher, C. R. Martin: Nature 393, 346 (1998)
82
[29] E. S. Steigerwalt, G. A. Deluga, D. E. Cliffel, C. M. Lukehart: J. Phys. Chem.
B 105, 8097 (2001) 82
[30] C. A. Bessel, K. Laubernds, N. M. Rodriguez, R. T. K. Baker: J. Phys. Chem.
B 105, 1115 (2001) 82
[31] M. Hirscher, M. Becher, M. Haluska, U. Dettlaff-Weglikowska, A. Quintel,
G. S. Duesberg, Y. M. Choi, P. Downes, M. Hulman, S. Roth, I. Stepanek,
P. Bernier: Appl. Phys. A-Mater. Sci. Process. 72, 129 (2001) 82
[32] L. Zhang, A. V. Melechko, V. I. Merkulov, M. A. Guillorn, M. L. Simpson,
D. H. Lowndes, M. J. Doktycz: Appl. Phys. Lett. 81, 135 (2002) 82
[33] J. Sandler, P. Werner, M. S. P. Shaffer, V. Demchuk, V. Altstadt, A. H. Windle:
Compos. Pt. A-Appl. Sci. Manuf. 33, 1033 (2002) 82
[34] E. Kymakis, G. A. J. Amaratunga: Appl. Phys. Lett. 80, 112 (2002) 82
[35] P. Calvert: Nature 399, 210 (1999) 82
[36] M. F. Yu, O. Lourie, M. J. Dyer, K. Moloni, T. F. Kelly, R. S. Ruoff: Science
287, 637 (2000) 82
[37] T. W. Ebbesen, P. M. Ajayan: Nature 358, 220 (1992) 82
[38] C. Journet, W. K. Maser, P. Bernier, A. Loiseau, M. L. delaChapelle,
S. Lefrant, P. Deniard, R. Lee, J. E. Fischer: Nature 388, 756 (1997) 82
[39] A. Thess, R. Lee, P. Nikolaev, H. J. Dai, P. Petit, J. Robert, C. H. Xu,
Y. H. Lee, S. G. Kim, A. G. Rinzler, D. T. Colbert, G. E. Scuseria, D. Tomanek,
J. E. Fischer, R. E. Smalley: Science 273, 483 (1996) 82
[40] K. B. T. Teo, M. Chhowalla, G. A. J. Amaratunga, W. I. Milne, G. Pirio,
P. Legagneux, F. Wyczisk, J. Olivier, D. Pribat: J. Vac. Sci. Technol. 20, 116
(2002) 83
[41] L. Delzeit, I. McAninch, B. A. Cruden, D. Hash, B. Chen, J. Han, M. Meyyap-
pan: J. Appl. Phys. 91, 6027 (2002) 83
[42] B. O. Boskovic, V. Stolojan, R. U. A. Khan, S. Haq, S. R. P. Silva: Nature
Mater. 27 Oct 2002 (2002) 83
[43] C. Bower, W. Zhu, S. Jin, O. Zhou: Appl. Phys. Lett. 77, 830 (2000) 83
[44] V. I. Merkulov, D. H. Lowndes, Y. Y. Wei, G. Eres, E. Voelkl: Appl. Phys.
Lett. 76, 3555 (2000) 83
92 Martin S. Bell et al.

[45] M. S. Bell, R. G. Lacerda, K. B. K. Teo, N. L. Rupesinghe, M. Chhowalla,


G. A. J. Amaratunga, W. I. Milne: Appl. Phys. Lett. 85, 1137 (2004) 84
[46] R. T. K. Baker, P. S. Harris: Chemistry and Physics of Carbon, vol. 14 (Dekker,
New York 1978) p. 83 84
[47] NIST Chemistry WebBook
URL http://webbook.nist.gov/ 85
[48] B. A. Cruden, A. M. Cassell, Q. Ye, M. Meyyappan: J. Appl. Phys. 94, 4070
(2003) 85
[49] M. Chhowalla, K. B. K. Teo, C. Ducati, N. L. Rupesinghe, G. A. J. Ama-
ratunga, A. C. Ferrari, D. Roy, J. Robertson, W. I. Milne: J. Appl. Phys. 90,
5308 (2001) 85, 86
[50] K. B. T. Teo, M. Chhowalla, G. A. J. Amaratunga, W. I. Milne, D. G. Hasko,
G. Pirio, P. Legagneux, F. Wyczisk, D. Pribat: Appl. Phys. Lett. 79, 1534
(2001) 85
[51] Y. S. Woo, D. K. Jeon, I. T. Han, N. S. Lee, J. E. Jung, J. M. Kim: Diam.
and Rel. Mats. 11, 59 (2002) 85, 87
[52] A. M. Cassell, S. Verma, L. Delzeit, M. Meyyappan: Langmuir 17, 266 (2001)
86
[53] L. Delzeit, C. V. Nguyen, R. M. Stevens, J. Han, M. Meyyappan: Nanotech-
nology 13, 280 (2002) 86
[54] Y. C. Choi, Y. H. Lee, B. S. Lee, G. Park, W. B. Choi, S. Lee, J. M. Kim: J.
Vac. Sci. Technol. A 18, 1864 (2000) 86
[55] V. I. Merkulov, D. H. Lowndes, Y. Y. Wei, G. Eres, E. Voelkl: Appl. Phys.
Lett. 75, 1086 (1999) 86
[56] Y. C. Choi, Y. M. Shin, S. C. Lim, D. J. Bae, Y. H. Lee, B. S. Lee, D. Chung:
J. Appl. Phys. 88, 4898 (2000) 86
[57] C. Bower, O. Zhou, W. Zhu, D. J. Werder, S. H. Jin: Appl. Phys. Lett. 77,
2767 (2000) 86
[58] R. T. Baker: Carbon 27, 315 (1989) 87
[59] A. R. Harutyunyan, T. Tokune, E. Mora: Appl. Phys. Lett. 86, 153113 (2005)
87
[60] J. B. O. Caughman, L. R. Baylor, M. A. Guillorn, V. I. Merkulov, D. H. Lown-
des, L. F. Allard: Appl. Phys. Lett. 83, 1207 (2003) 87
[61] D. Hash, D. Bose, T. R. Govindan, M. Meyyappan: J. Appl. Phys. 93, 6284
(2003) 87
[62] M. Jung, K. Y. Eun, J.-K. Lee, Y.-J. Baik, K.-R. Lee, J. W. Park: Diam. and
Rel. Mats. 10, 1235 (2001) 88
[63] D. B. Hash, M. S. Bell, K. B. K. Teo, B. A. Cruden, W. I. Milne, M. Meyyap-
pan: Nanotechnology 16, 925 (2005) 89

Index
a-C, 83 C60 , 78, 79
C70 , 79
arc discharge, 82
catalytic chemical vapour deposition
buckminsterfullerene, 78, 79 (CVD), 83
Growth Mechanism During PECVD of Multiwalled Carbon Nanotubes 93

chemical vapour deposition (CVD), 83 chiral vector, 79


chiral vector, 79 growth mechanism, 85
multiwalled nanotubes (MWNTs),
glow discharge, 84 77, 79, 82, 84
graphene, 78–80 single-walled nanotubes (SWNTs),
79, 81, 82
laser vaporisation, 82
zig-zag nanotubes, 79, 80
mass spectrometry, 84, 85
plasma, 83
nanofibres, 80, 82 plasma-enhanced chemical vapour
herringbone nanofibres, 81 deposition (PECVD), 77, 83, 84,
stacked nanofibres, 81 87, 89
nanotubes, 77–85, 87–89
armchair nanotubes, 79, 80 residual gas analysis (RGA), 84, 85, 88,
chiral nanotubes, 79 89

Anda mungkin juga menyukai