1-S2.0-S0008622309002036-Growth Time

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

available at www.sciencedirect.com

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

Dimensional control of multi-walled carbon nanotubes


in floating-catalyst CVD synthesis

Gregg S.B. McKeea, Christian P. Decka, Kenneth S. Vecchiob,*


a
Materials Science and Engineering Program, University of California-San Diego, La Jolla, CA 92093, USA
b
Department of NanoEngineering, University of California-San Diego, 9500 Gilman Drive, MC-0448, La Jolla, CA 92093, USA

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

Article history: Dimensional control in CVD synthesis of MWNT’s is significant and critical to a number of
Received 15 July 2008 different applications. This study examines the dimensional effect of a number of synthe-
Accepted 27 March 2009 sis variables on the products of floating-catalyst CVD, including catalyst concentration in
Available online 5 April 2009 the feedstock, nanotube growth time, and deposition substrate selection. Extensive diam-
eter surveys are performed by TEM and compared with results from thermo-gravimetric
analyses to Raman spectroscopy, offering a novel dimensional analysis of nanotubes grown
by FC-CVD methods. CNT diameters are inversely proportional to the catalyst concentra-
tion with weak correlation over the range examined and are directly proportional to growth
time. Results are combined with prior art to develop a new theory regarding catalyst parti-
cle formation over a range of catalyst concentrations. Carbon deposition occurs in two
stages, the first characterized by accelerating deposition and increases in CNT diameter
and length, the second by etching of the array and carbon deposition at a constant rate.
Deposition substrates interact directly with the catalysts to strongly influence the resulting
nanotube diameters, based upon the mobility of the catalysts on the substrate surface.
 2009 Elsevier Ltd. All rights reserved.

1. Introduction amount of literature and the wide variety of synthesis varia-


tions available for CNT’s.
Given the extraordinary properties of carbon nanotubes Floating-catalyst chemical vapor deposition (FC-CVD) syn-
(CNT’s), from high specific surface area to tunable electric thesis of CNT’s has its origins in the synthesis of vapor-grown
conductivity to high specific stiffness and strength, there carbon fibers (VGCF). For many years CNT’s were observed in
has been a rush to develop applications for nanostructures VGCF products, but not recognized as anything but particu-
before the methods for synthesis are adequately understood. larly small fibers. In recent years, reports have suggested at
Field emission devices [1–3], nano-electric devices [1,4–6], and least some VGCF originate from or otherwise contain hollow
structural composites [7,8] are emerging, making use of CNT’s CNT cores [12,13]. The synthesis process and proposed
as core elements. However, the synthesis of economical, growth models for the two are nearly identical [14–16]. For
reproducible, quality nanotubes remains a significant chal- this reason, it may be expected that some of the synthesis
lenge, which must be addressed before widespread adoption factors influencing the properties of resulting VGCF might
of carbon-nanotube or any nanostructure-centric application also be significant in the synthesis of CNT’s. For example,
can occur. To date, studies in CNT dimensional control re- the ratio of catalyst to carbon source has been shown to be
main very unusual [9–11], especially given the enormous proportional to the fiber diameter and aspect ratio of VGCF
[16]. If the same were applicable to multi-walled nanotubes

* Corresponding author: Fax: +1 858 534 5698.


E-mail address: kvecchio@ucsd.edu (K.S. Vecchio).
0008-6223/$ - see front matter  2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.carbon.2009.03.060
2086 CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

(MWNT), it would allow for more precise control over dimen- growth time are identified. An analysis of the substrate choice
sions in an area where most dimensional control is per- highlights newly observed interactions between the substrate
formed empirically, if at all. FC-CVD is quickly emerging as and catalysts delivered by the FC-CVD method.
one of the most scalable and least expensive methods for pro-
ducing high quality single-walled nanotubes (SWNT) and 2. Experimental
MWNT, thus, there is great need for a better understanding
of the variables, which define the resulting products. CNT’s were synthesized through the pyrolytic decomposition
Numerous studies have examined the importance of dif- of a benzene-ferrocene feedstock solution in a CVD furnace as
ferent growth parameters in catalyst assisted CVD growth of shown in Fig. 1. Nanotube synthesis began by flowing 99.5%
carbon nanotubes; however, these reports tend to focus argon through the reaction chamber (a 2’’ diameter quartz
exclusively on synthesis methods using thin film catalysts, tube) as the furnace heated to the synthesis temperature.
and the control of nanotube diameters in floating-catalyst Once the furnace reached the desired temperature (approxi-
CVD has not been widely researched. With thin film catalysts, mately 750 C at the growth zone, 120 C at the feedstock
nanotube diameters are controlled by varying the metal film injection point), the ferrocene–benzene feedstock solution
deposition thickness [17], but this approach cannot be applied was fed into the chamber, where it was collected at the
to FC-CVD, as the method of catalyst introduction is funda- mouth of the tube in an alumina crucible, approximately
mentally different. Instead of thin films, which coalesce into 300 mm from the growth zone, and vaporized by heat from
catalyst islands during heating, catalyst material is continu- the furnace. About 100 sccm Ar (99.5%) flow carried the vapor
ously added during floating-catalyst growth, and catalyst par- into the furnace, where CNT deposited onto substrates over a
ticles form as reactants decompose and deposit on substrate length of 75 mm. After the desired growth time elapsed, any
surfaces. As a result, nanotube diameters are affected by remaining solution was removed, the furnace shut down, and
parameters that do not apply to conventional thin film CVD argon flow maintained until the chamber had completely
growth, such as catalyst solution concentrations. Of the few cooled.
reports on diameter control using FC-CVD, the conclusions The catalyst concentration, growth time, and substrate
with regard to catalyst concentrations have not always been type were examined separately. In each set of experiments,
consistent, with increasing concentrations resulting in either all variables not under examination (furnace temperature,
reduced diameters [18], increased diameters [19–21], or show- growth time, substrate, etc.) were held approximately
ing little influence [22,23]. Of these reports, some examine constant.
only single-walled tubes [21,23], and those studying multi- In order to examine the significance of catalyst concentra-
walled tubes use widely different reaction temperatures tion, the ratio of ferrocene to benzene in the feedstock was
(ranging from 675 C [19] to 1150 C [18]), and catalyst concen- systematically varied, from 4.75 g ferrocene per liter benzene
trations (ferrocene-hydrocarbon molar ratios of 1.17–15.1% (0.227% molar ratio) up to the room temperature solubility
[18], 0.6–6.0% [19], 0.10–5.26% [20], and 0.44–1.73% [22]). Con- limit at approximately 200 g ferrocene per liter benzene
trol of nanotube diameters is important, and the mechanisms (9.58% molar ratio). Multiple synthesis runs were made for
by which the catalyst concentration can affect nanotube each catalyst to carbon ratio for purposes of analysis
dimensions must be understood. To further investigate this duplication.
relation, an intermediate reaction temperature (750 C) and Growth time was varied from 5 min up to 240 min by con-
catalyst concentration range (0.23–9.56% molar ratio) were se- trolling the time over which the catalyst/ferrocene feedstock
lected for this work. was introduced into the furnace (at constant rate). A variety
This work examines several variables relevant to dimen- of substrates were examined: silicon, quartz, alumina, NiO,
sional control in vapor-grown carbon fiber synthesis and their mullite, graphite, an alumina-mullite mixture, a machinable
dimensional impact on FC-CVD synthesis of MWNT. We pro- ceramic with approximate composition of 55% fluorophlogo-
pose herein a mechanism to explain these apparently con- pite mica in a borosilicate glass matrix, and a soda-lime glass
tradicting reports on the influence of catalyst (72.2% SiO2, 14.3% Na2O, 1.2% K2O, 6.4% CaO, 4.3% MgO, 1.2%
concentrations on nanotube diameter. At lower catalyst con-
centrations (such as those used by Sinnott et al. [19] and
Singh et al. [20]) nanotube diameters increase with increasing
catalyst concentrations, while at high concentrations (used
by Bai et al. [18]), larger metal particles form, which are not
suitable for nanotube formation. In addition, other reports
have indicated a relationship between substrate material
and nanotube diameter for thin film catalyst CVD growth
[24]. As most previous parametric studies with FC-CVD used
only a single substrate material (most often quartz), different
substrates were selected in this work to see if this parameter
had any effect on diameters with FC-CVD. Fig. 1 – Schematic diagram of the FC-CVD CNT growth setup.
Analysis is accomplished through extensive TEM diameter A solution of the catalyst (ferrocene) and carbon source (ben-
surveys, as well as Raman, thermo-gravimetric, and atomic zene) is introduced into the furnace in liquid form, thermally
force microscopy (AFM) analyses. Novel structure, reactivity, vaporized, and carried over the substrate where aligned CNT
and dimensional trade-offs with catalyst concentration and arrays are deposited.
CARBON 4 7 ( 20 0 9 ) 2 0 8 5–20 9 4 2087

Al2O3, 0.3% SO3). All substrates were polished to at least 3. Results and discussion
4000 grit prior to CNT deposition. Growth time and sub-
strate-related experiments used a catalyst to carbon ratio of CNT were deposited in aligned arrays with thicknesses from
approximately 15 g/L. To measure catalyst mobility on sub- 100 lm to a few millimeters (Fig. 2). Each sample was tested
strates, 10 nm of Fe was e-beam evaporated upon the sub- as-synthesized (excepting mechanical processing for TGA &
strates in question, which were then sintered for 10 min sonication for TEM sample prep) without treatment for the re-
under approximately 150 ccm of 10% H2 90% Ar, to replicate moval of non-nanotube impurities. Samples were observed to
the reducing atmosphere present in the synthesis chamber be generally free of non-nanotube carbonaceous impurities at
during nanotube growth. SEM levels of magnification. At higher magnifications in the
Residual iron content in as-produced nanotubes was cal- TEM; however, certain samples were observed to have poten-
culated through complete oxidative removal of the carbona- tially significant levels of carbonaceous and catalyst particle
ceous component. Large samples (5–100+ mg) were held at impurities, as will be discussed later.
650 C in air for several hours and the total mass change re- For the purposes of this investigation, cleaning of the CNT
corded. By assuming complete conversion of the catalyst to by refluxing or another partial oxidation technique may have
Fe2O3, each residual mass was corrected for oxygen mass removed excess catalyst particles and non-nanotube carbo-
(Oxygen makes up 30.06% of Fe2O3 mass) and the mass of re- naceous impurities, but would have resulted in significant
tained iron calculated. damage to the CNT structures, such that any meaningful
Deposition rates were calculated from the amount of car- structural analysis would have been impossible.
bon deposited on a 75 · 25 mm quartz slide placed at the As might be expected, a strong correlation between the
beginning of the growth zone in the reactor. catalyst to carbon ratio and the retained iron content is ob-
Nanotube dimension measurements were made via SEM served (Fig. 3). The relationship roughly follows a parabolic
(Philips XL20, 15 kV accelerating voltage) or TEM (Philips curve, with a rapid increase in the retained iron content at
EM420, 120 kV accelerating voltage, or JEOL JEM-2011, 200 kV low catalyst concentrations (0–25 g/L), followed by a more
accelerating voltage) observations. TEM samples were pre- gradual increase at higher concentrations. Some variability
pared through sonication in MeOH for 2.5 min and dripped in iron content is evident for synthesis runs at a given catalyst
onto a 200 mesh Cu grid. Micrographs were collected between concentration, but the overall pattern shows a proportional
70,000 and 150,000· magnification and nanotube diameters in increase in one with the other. In only one of the many syn-
each measured and tabulated. At least 250 individual and dis- thesis runs did the retained iron content drop below
tinct nanotubes from each sample were measured to provide 1.25 wt.% Fe, possibly suggesting a minimum iron content
a statistically accurate representation. Catalyst particle size for CNT formation (catalyst limited growth). At low catalyst
measurements were made by contact-mode atomic force concentrations, a situation of excess carbon may exist (cata-
microscopy (Nano-R, Pacific Nanotechnologies). lyst limited growth), such that amounts of carbon cannot be
Oxidation mass-loss profiles were obtained in a Perkin–El- incorporated into the CNT and may deposit later in the fur-
mer Pyris Diamond Thermo-Gravimetric/Differential Thermal nace, beyond the CNT deposition area, or be carried out in
Analyzer (TG/DTA). Masses of approximately 1 mg were the exhaust stream. Saturation of the retained iron content
placed in open-top alumina pans under a purge gas of 2% with high catalyst concentrations may imply an upper limit
O2, balance He. The thermal program initiated the purge at which point there is insufficient carbon for the formation
gas, raised the temperature to 200 C at 40/min, soaked for of a nanotube from each catalyst particle (carbon limited
10 min, then increased to 1000 C at the rate of 1/min (in or- growth).
der to simulate the reaction equilibrium state and allow direct
comparison between these data and those found in other
works) [25–27]. Each sample was analyzed twice to ensure
repeatability of the results.
Raman spectroscopy (Renishaw Raman) was performed at
room temperature using an argon ion laser (514.5 nm,
0.35 mW @ sample). Data was collected from three distinct
locations of the exposed side of an aligned array of nanotubes
for each sample. Three spectra collected for each location were
normalized to the D-band (1350 cm 1) peak and averaged to-
gether to eliminate minor variations in the individual spectra.
In order to examine the relationship between catalyst con-
centrations, carbon levels in the furnace, and nanotube diam-
eters, a two-stage CVD synthesis was performed with the goal
of creating an environment of excess carbon for the cases of
high and low catalyst concentration. In the first stage of syn-
thesis, a growth feedstock with either 12.5 or 151 g ferrocene/ Fig. 2 – An example SEM micrograph showing the macrosco-
L benzene was used. After 2.5 min of growth, the feedstock pic alignment of the CNT array subsequent to removal from
containing ferrocene was removed and replaced with pure the growth substrate. The degree of alignment may be con-
benzene for an additional 27.5 min. Dimensions of the result- sidered generally representative of all aligned arrays exami-
ing deposits were then found via TEM. ned. Scale bar is 200 lm.
2088 CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

in the data suggest no dependence of the diameter on the iron


content. However, if one chooses to exempt the 12 wt.% Fe
data point, a trend is weakly suggested, where the mean
diameter shows a slight decreasing trend with increasing iron
content. In either case, no trend is evident in the diameter’s
standard deviation. The diameter trend is in agreement with
previous results from Bai et al. [28], who studied catalyst con-
centrations from 1.17 to 15.1 at.%, but contrary to results by
Sinnott et al. [29], who reported on catalyst concentrations
of 0.075 at.% and 0.75 at.%. However, it should be noted that
the latter reported on only two samples/catalyst
concentrations.
Bai et al. [28] suggest that the increased catalyst content
creates a finer dispersion of catalyst particles as well as a car-
Fig. 3 – Retained iron content plotted as a function of the fer- bon deficiency in the reaction zone, resulting in a smaller
rocene to benzene ratio in the synthesis feedstock. The cur- mean CNT diameter. In order to test this hypothesis, a two-
ve shows a direct proportionality between the two variables stage CVD synthesis was performed with the goal of creating
and is roughly parabolic up to the room temperature solubi - an environment of excess carbon for the cases of high and
lity limit of ferrocene in benzene (200 g/L). low catalyst concentration, as discussed in Section 2. CNT
dimensions resulting from these runs fell generally in line
with the other diameter measurements shown in Fig. 5: the
Altering the catalyst concentration in the feedstock with-
12.5 g/L sample showed a mean diameter of 37.2 nm and a
out modifying other synthesis conditions does not appear to
standard deviation of 15.3 nm; the 151 g/L sample showed a
directly change the resulting deposition rate (Fig. 4). Large
mean diameter of 33.8 nm and a standard deviation of
changes in the catalyst concentration produce no predictable
19.3 nm. Histograms showing the numerical distribution of
change in the yield, when performed under similar circum-
the two samples may be found in the Supplementary
stances. Similarly, a large range of deposition rates can be ob-
information.
tained using very similar concentrations (Fig. 4, note yield
The observed differences in standard deviation fall within
range at 15 g/L) by varying the growth time. Data points at
the scatter observed for all catalyst concentrations and can-
concentrations other than 15 g/L had growth times of approx-
not be considered significant. The difference in diameters is
imately 90 min. Thus, it is apparent that the relationship be-
approximately 10% of the overall diameter and as such, falls
tween yield and ferrocene concentration, if any, is minimal in
above the approximate margin of error of 6% for a 95% confi-
comparison with the relationship between yield and growth
dence level. If the theory put forward by Bai et al. [28] were
time.
accurate, the two samples should have statistically similar
Results of the survey of diameter distribution as a function
diameters, but, as a difference has been observed, it must
of retained iron content are summarized in Fig. 5. More de-
tailed distribution results for individual samples may be
found in the Supplementary information. Significant scatter

Fig. 5 – CNT diameter and standard deviation are plotted


Fig. 4 – Deposition rate (yield) is plotted as a function of versus the retained iron content. A weak decreasing trend is
catalyst concentration. Wide variations of the catalyst con- observed in the mean diameter as retained iron content in-
centration produce no predictable change in yield. However, creases. Significant scatter is apparent in the standard dev-
by varying the growth time, large changes in yield may be iation and no direct relation to the retained iron content is
obtained without altering the catalyst concentration. Thus, observed. Data points represent a simple random sample of
yield appears to be independent of the iron to carbon feed- at least 250 data points, excepting the highest iron content
stock ratio, but is strongly dependent upon growth time. sample (12 wt.% Fe), for which 206 tubes were examined.
CARBON 4 7 ( 20 0 9 ) 2 0 8 5–20 9 4 2089

be due to causes other than a carbon deficiency during


synthesis.
It would be generally expected that a higher catalyst con-
centration will lead to more and larger catalyst particles than
a comparatively lower catalyst concentration. It is also ex-
pected that the growth rate of the catalyst particles (the rate
at which they will encounter reactant moieties in the furnace,
incorporate them, and increase the particle’s dimensions) will
depend directly upon the dimensions of the particle at the
time in question. Once a catalyst particle’s dimension sur-
passes some threshold; however, the likelihood that a CNT
will grow from it diminishes, even as the particle’s growth
rate continues to increase. Thus, the nanotubes that form will
do so from smaller catalyst particles (on average) with com-
mensurately smaller diameters. The larger catalyst particles
will tend towards faster growth to the point at which they ex-
clude themselves due to size, leading to a smaller average
CNT diameter with the higher catalyst concentrations. On
the other hand, at very low concentrations (lower than those
explored here), the particles would be size-limited due to an
unavailability of catalyst material, leading to a proportional
relationship between catalyst concentration and CNT
diameters.
When viewed in conjunction with the two previously men-
tioned reports [28,29], a more generalized picture may be ob-
served: at very low catalyst concentrations, catalyst particle
size (and thus, CNT diameter) is directly proportional to the
catalyst concentration. As the catalyst concentration is in-
Fig. 6 – Raman studies of CNT structure are summarized in
creased; however, many catalyst particles grow to exceed
terms of the (a) G:D and (b) G’:D peak ratios, plotted versus
the threshold for CNT formation, leading to increased non-
the retained iron content in the sample. The same approx-
CNT contamination in the sample and the formation of smal-
imate minimum of peak ratios is observed across all iron
ler diameter nanotubes.
contents; however, as iron content is decreased, the maxi-
There exists considerable scatter in the mean CNT diame-
ma of the data range increases as well. This suggests the
ters for all catalyst concentrations investigated. As such,
potential, but not guarantee, for improved structures with
while the catalyst concentrations examined herein may im-
decreasing iron contents.
pact the CNT dimensions, variation within this range does
not appear to be a reliable method for controlling their
dimensions, as it is in the case of VGCF’s. Other variables ap-
pear to have significantly larger and more reliable impacts. the data range decreases, with approximately the same min-
Raman spectroscopy performed on MWNT’s generally re- imum for all iron contents measured. This suggests that
sults in a characteristic fingerprint: the D band, which ap- CNT’s with lower iron content have the potential for signifi-
pears at approximately 1350 cm 1, is associated with sp3- cantly improved structural perfection versus those with larger
hybridized carbon and disordering in the nanotube, while amounts of retained iron, but are not guaranteed to be so im-
the G band appears at approximately 1580 cm 1, a result of proved. This also contradicts the finding by Bai et al. [28] that
sp2-hybridized carbon bonding (the in-plane stretching mode impurity content in CVD generated CNT’s decreases when the
for graphene sheets) [30]. The G’ band is the second order har- catalyst to carbon source ratio is increased. However, that
monic of the D band and appears in the realm of 2700 cm 1. study appears to have been performed by SEM observations,
Other higher order peaks occur in the neighborhood of the so results may have been more representative of localized
G’ peak as well, but are generally not useful for CNT charac- areas rather than bulk materials.
terization. As CNT quality measurements are frequently per- While the as-synthesized CNT arrays were observed to be
formed through comparison of the G to D band peaks in the clean on the micron scale (Fig. 7a), nanoscale impurities were
Raman spectra [31,32], Fig. 6a summarizes the Raman spec- observed in certain samples. For example, samples using cat-
troscopy results in this fashion. However, it has been shown alyst concentrations equal to or greater than approximately
that the G-band resonance in CNT’s shows some defect sensi- 75 g ferrocene per liter benzene were often found to contain
tivity and thus, comparison of the G’ band peak (2700 cm 1) significant quantities of carbon coated catalyst particles, both
to the D band peak may be a more accurate measure of CNT isolated and attached to nanotubes, in addition to those ob-
quality [33]; G’:D peak ratios are presented in Fig. 6b. served enclosed within CNT’s (Fig. 7b). As this type of contam-
While the Raman spectroscopy results summarized in ination was generally limited to the higher catalyst ratios
Fig. 6a and b shows a wide range of G:D and G’:D peak ratios used and not observed in other samples, it is likely a direct re-
at low retained iron content, as the iron content is increased, sult of the higher concentrations used in synthesis, and corre-
2090 CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

Fig. 8 – Thermo-gravimetric burn-off of CNT samples is sum


marized and plotted versus the sample iron content. In ord-
er to summarize the oxidation curve with a single number,
the temperature corresponding to oxidation of half the sam-
ple mass is used. As iron content in the sample decrea-ses,
burn-off temperatures increase, indicating less reactive str-
uctures. This change occurs through diminished defect den-
sities and less catalytic oxidation by the iron particles.

The second mechanism arises from the improved struc-


tural perfection demonstrated in the Raman spectroscopy
analyses reported here. Defects in CNT’s serve as sites of in-
creased reactivity [35,38], so as the peak ratio increases, indi-
cating a smaller density of defect structures, the reactivity of
Fig. 7 – (a) An example SEM micrograph of a CNT sample. the CNT structure will decrease. Both mechanisms serve to
The CNT’s examined appear free of impurities on the mic- decrease the reactivity of the CNT’s and thus, increase the
ron scale. Scale bar for this micrograph is 2 lm. (b) Samples sample’s temperature of oxidation.
synthesized using the higher catalyst concentrations analy- Elongation of the CNT array and carbon deposition rate ap-
zed were frequently observed to contain significant quanti pear to exhibit two distinct stages as growth time is increased
ties of carbon coated catalyst particles in addition to those (Fig. 9a and b). Throughout the first stage (stage 1 growth), ar-
observed enclosed within CNT’s. As this type of contamina- ray length increases linearly and the deposition rate acceler-
tion was not observed at lower catalyst concentrations, this ates (as noted by the increasing slope of the fit line in
is likely a direct result of higher catalyst concentrations. Sca- Fig. 9b). In the second stage (stage 2 growth), array length be-
le bar is 100 nm. gins to diminish while the mass continues to increase, but
now at a linear rate.
Deposition and growth during stage 1 may be understood
lates well with the proposed catalyst particle formation if one considers the changing morphology of the deposition
scheme. surface. Initially, only the deposition substrate is present. At
A summary of thermo-gravimetric oxidation analyses ap- this point, carbon is deposited onto the substrate surface,
pears in Fig. 8. As the mass-loss curves are asymmetric over generally as nascent CNT’s, or is incorporated into their elon-
temperature and in order to concisely compare oxidation pro- gation. As the tubes elongate, wall area is increased, thereby
files for many samples, the data points plotted in Fig. 8 corre- increasing the available surface area for carbon deposition.
spond to a reaction coordinate of x = 0.5 (each profile showed Carbon continues to be incorporated into the elongation of
only a single oxidation peak indicative of CNT burn-off). The the CNT’s, but secondary hydrocarbon pyrolysis leads to car-
oxidation profiles reveal an increase in the oxidation temper- bon deposition on the sidewalls, increasing CNT diameters
ature with decreasing iron content. This indicates a decreased (and further increasing the wall area available for deposition).
reactivity in the CNT, which stems from at least two different Diameter survey results (Fig. 10) confirm an increase in CNT
mechanisms. diameter and standard deviation with increasing growth
The first arises from the presence of iron particles catalyz- time. Thus, as the available surface area for deposition in-
ing the oxidation of CNT’s [34–37]. As the amount of retained creases, so the rate of deposition increases. The increase in
iron increases, the amount of catalytic material is increased, the standard deviation with growth time may be attributed
resulting in a higher density of catalytically active sites. As to the constant nucleation of nanotubes. As growth time pro-
the number of sites is diminished (through reduction of the gresses, carbon deposition will occur on the surfaces of exist-
iron content), the reactivity of the structure decreases, shown ing nanotubes, increasing their diameters, but will also
by an increase in the temperature of oxidation. contribute to the formation of new tubes which will grow in
CARBON 4 7 ( 20 0 9 ) 2 0 8 5–20 9 4 2091

new catalysts can no longer reach the substrate and base of


the array due to CNT thickening or (2) all active catalyst par-
ticles have become poisoned and can no longer support
CNT growth. The decrease in the average array length in con-
cert with the increasing carbon mass also suggest preferential
etching of the top (more exposed) surface of the array and
redistribution or redeposition of that carbon into the bulk of
the array, leading to a densification of the material. Addition-
ally, carbon continues to deposit, but does not appear to be
incorporated into the tubule structures. Thus, the available

Fig. 9 – (a) CNT array length and (b) deposition rate are plot-
ted as a function of growth time. Both may be separated into
two phases of growth and deposition. Stage 1 growth shows
linear array elongation and accelerating deposition rate
with increasing time. Stage 2 growth shows etching of the
CNT array and continued carbon deposition, but now at a li-
near rate.

Fig. 11 – (a) Nanoscale non-nanotube carbonaceous conta-


Fig. 10 – CNT diameter and standard deviation are plotted
mination was observed on larger CNT’s in several samples.
versus the growth time during stage 1 growth. Accelerating
This type of contamination occurred in samples with incr-
carbon deposition during stage 1 occurs as the surface area
eased growth times, across the spectrum of catalyst concen-
available for deposition increases as CNT’s elongate. This
trations examined. Scale bar is 100 nm. (b) A bi-layer struc
results in an increase in average CNT diameters as well as in
ture was observed on some of the CNT’s examined, causing
the diameter’s standard deviation as growth time increases.
significant variation in the outer diameter of the CNT, while
the inner diameter and inner layer thickness remained app-
roximately constant. Scale bar is 100 nm. (c) The inner layer
diameter with passing time, thereby increasing the overall appears to be organized graphene layers, while the outer
distribution of CNT diameters with growth time. layer appears to be a less organized form of carbon, proba-
Growth changes to stage 2 at approximately 120 min. Two bly amorphous. White lines indicate the approximate boun-
potential reasons behind this change include: (1) carbon and daries of each layer. Scale bar is 20 nm.
2092 CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

ness, while the inner layer’s thickness and the tube’s inner
diameter remain approximately constant (Fig. 11b). In these
cases, the tube structure appears to be an inner set of orga-
nized graphene layers coated with a less organized outer layer
of carbon, probably amorphous (Fig. 11c), supporting the pro-
posed mechanism for stage 1 growth.
Fig. 12 summarizes the results of the diameter survey on
the various substrates tested. No deposits were observed on
graphite substrates and the NiO, soda-lime glass, and alu-
mina/mullite combination substrates did not yield well-
aligned CNT arrays. Notably, a significant difference in mean
CNT diameter is observed between substrate materials. Most
substrates tested yielded mean CNT diameters of approxi-
mately 25 nm. Alumina and quartz substrates fell signifi-
cantly outside this range, with mean diameters of 15.8 nm
and 36.1 nm, respectively.
It has been recently suggested that iron mobility may vary
Fig. 12 – Mean CNT diameter and standard deviation are
significantly from one substrate material to another [39], lead-
compared across several different deposition substrates.
ing to the diameter variations observed. A lower catalyst
Substrate impact on CNT dimensions appears to be related
mobility on the substrate would limit sintering and agglomer-
to the mobility of the catalyst particles on the substrate
ation of the catalyst, thereby yielding smaller catalyst parti-
surface, which governs the ability of the catalysts to sinter
cles and a lower mean CNT diameter. Given the CNT
and agglomerate.
diameter differences observed, it is expected that iron mobil-
ity on alumina would be lower than on silicon, which, in turn,
surface area does not significantly change as a result of con- would be lower than on quartz. Examination of 10 nm iron
tinued deposition, yielding a linear increase in mass and con- thin-films deposited on these substrates and sintered for
stant deposition rate. For these reasons, with increasing 10 min confirms this prediction (Fig. 13). Particles are modeled
growth times, synthesis products transition from aligned ar- as semi-ellipsoid shapes and increase in size from alumina
rays of CNT to aligned arrays of nano-fibers to a CNT-rein- (radii: 275, 271, and 31 nm) to silicon (radii: 315, 280, and
forced-carbon composite material. 45 nm) to silica (radii: 429, 364, and 69 nm).
Nanoscale carbonaceous contamination was observed on It is proposed that the observed diffusivity difference be-
larger tubes in certain samples (Fig. 11a). In this case, there tween silica and alumina is related to the inter-atomic bond-
is no obvious link between the catalyst concentration ratio ing within the oxides which influences the ease with which
used and the contaminated samples; several samples across catalyst oxides may form as well as the interactions of the
the range of ratios examined displayed varying degrees of car- catalyst oxide with the substrate. The bond character in alu-
bonaceous deposits attached to the CNT walls. Instead, this is mina tends to be very ionic [40], whereas the bond character
a result of longer growth times, showing the impact of addi- in silica tends to be mostly covalent [41]. Formation of iron
tional carbon deposition on CNT walls. Additionally, some oxides, also predominantly ionic, has been observed in FC-
samples with this type of contamination showed an interest- CVD on both silica and alumina substrates [42]. The iron
ing bi-layer structure on certain CNT’s, similar to that noted oxide may undergo ionic interactions with the alumina to
elsewhere [20,31]. The outer layer varies significantly in thick- form a solid solution or potentially a new compound at the

Fig. 13 – About 5 · 5 lm AFM scans of (a) Alumina, (b) Silicon, and (c) Quartz substrates. About 10 nm of iron was e-beam
evaporated onto each of the substrates, which were then annealed at 850 C for 10 min under 10% H2, 90% Ar gas stream.
Resulting particle volumes increase, left to right, indicating increasing mobility of the catalyst on the substrate surface and
explaining the difference in CNT dimensions observed. Particles are modeled as semi-ellipsoid shapes; alumina: radii: 275,
271, and 31 nm; silicon: radii: 315, 280, and 45 nm; silica: radii: 429, 364, and 69 nm. Scale bars are 1 lm.
CARBON 4 7 ( 20 0 9 ) 2 0 8 5–20 9 4 2093

substrate surface (spinel structured Fe(Fe,Al)2O4) [43], pinning Substrate choice has been newly shown to have significant
the particle and thus, limiting its mobility on the surface. Fur- impact upon CNT diameters grown by FC-CVD, with the rela-
ther, above 600 C amorphous (sputtered) SiO2 surfaces trans- tive effect predictable through the catalyst’s mobility upon
form to non-polar 4-membered siloxane rings [44], and would the substrate’s surface. In the case of alumina, it is proposed
not have similar interactions with the iron or iron oxide, that catalyst mobility is limited by ionic substrate/catalyst or
although an olivine structure with comparatively higher cata- substrate/catalyst-oxide interactions. These results should al-
lyst content may be formed. Thus, the use of ionically bonded low for improved control in the FC-CVD synthesis of CNT’s.
substrates, particularly oxides with which the catalyst oxide
may form a solid solution, should favor smaller catalyst for- Acknowledgements
mation during the sintering phase. This topic will be exam-
ined in more detail in a later paper. The authors acknowledge Dr. Ron Gronsky and Dr. Chris Ku-
mai of UC Berkeley, and Dr. Frank Talke and Ralf Brunner of
4. Conclusions UC San Diego CMRR for assistance accessing various pieces
of analytical equipment.
Given the similarity of VGCF synthesis to FC-CVD CNT synthe-
sis, it was expected that CNT diameter would similarly depend Appendix A. Supplementary information
upon the catalyst to carbon source ratio. Catalyst particle
dimensions are traditionally understood to directly impact Supplementary data associated with this article can be found,
the diameters of the CNT’s formed; however, in this case, a in the online version, at doi:10.1016/j.carbon.2009.03.060.
higher catalyst concentration in the feedstock does not directly
translate to larger CNT dimensions. The results of the TEM
diameter survey indicate a relatively weak correlation between R E F E R E N C E S
CNT dimensions and the catalyst concentrations examined,
but with the opposite proportionality versus what was ex-
pected. It has been shown that this dependence is not a result [1] Sveningsson M, Lee SW, Park YW, Campbell EEB. Field
of carbon depletion within the reaction zone. Instead, the new- emission from multi-walled carbon nanotubes – its
ly proposed theory suggests that at very low catalyst concen- applications in NEMS. AIP Conference 2005:620–3.
trations, catalyst particle size is limited by the availability of [2] Choi WB, Chung DS, Kang JH, Kim HY, Jin YW, Han IT, et al.
catalyst material and thus, directly proportional to the catalyst Fully sealed, high-brightness carbon-nanotube field-
emission display. Appl Phys Lett 1999;75(20):3129–31.
concentration, but as the concentration is increased, catalyst
[3] de Heer WA, Chatelain A, Ugarte D. A carbon nanotube field-
growth accelerates proportionately with size and many parti-
emission electron source. Science 1995;270(5239):1179–80
cles grow to exceed the threshold for CNT formation, leading (Washington, D.C.).
to increased non-CNT contamination in the sample and CNT [4] Bandaru PR, Daraio C, Jin S, Rao AM. Novel electrical
formation from the smaller catalysts. switching behaviour and logic in carbon nanotube Y-
It is evident that there are trade-offs for catalyst concen- junctions. Nat Mater 2005;4(9):663–6.
tration and growth time. As catalyst concentration increases, [5] Li H, Zhang Q, Li J. Carbon-nanotube-based single-electron/
hole transistors. Appl Phys Lett 2006;88(1):013508/1–3.
diameter decreases, but stability and structural perfection de-
[6] Johnston DE, Islam MF, Yodh AG, Johnson AT. Electronic
crease, raising the reactivity of the CNT’s. As growth time in-
devices based on purified carbon nanotubes grown by high-
creases, the overall CNT array length increases, but so do the pressure decomposition of carbon monoxide. Nat Mater
average diameter and its standard deviation. Carbon deposi- 2005;4(8):589–92.
tion rate appears to be independent of the catalyst concentra- [7] Kireitseu MV, Bochkareva LV, Hui D. Rheological behavior of
tion, but increases in two stages with the growth time, with CNT-reinforced damping materials. Ann Trans Nordic Rheol
CNT yielded only during the first growth stage. The examina- Soc 2006;14:219–25.
[8] Hernandez CD, Zhang M, Fang S, Baughman RH, Gates TS,
tion of FC-CVD in this work revealed a two-stage growth pro-
Kaing SK. Multifunctional characteristics of carbon nanotube
cess, which is unique to floating-catalyst synthesis. With thin (CNT) yarn composites. In: Proceedings of the
film catalysts, an initial rapid growth slows as the catalysts multifunctional nanocomposites international conference,
become poisoned. This process can occur relatively quickly Honolulu, HI, United States. 2006. p. mn2006 17028/1–
(<1 h), and no substantial changes to nanotube lengths or mn2006/5.
yields are observed beyond this point [45,46]. The continuous [9] Kuo C-S, Bai A, Huang C-M, Li Y-Y, Hu C-C, Chen C-C.
Diameter control of multiwalled carbon nanotubes using
addition of catalyst material in FC-CVD allows extended reac-
experimental strategies. Carbon 2005;43(13):2760–8.
tion durations, and in this study a second growth stage was
[10] Kim L, Lee E-M, Cho S-J, Suh JS. Diameter control of carbon
identified (beyond some initial growth at a constant rate). In nanotubes by changing the concentration of catalytic metal
this stage the nanotube mat stops lengthening (and actually ion solutions. Carbon 2005;43(7):1453–9.
shrinks as products are etched away); however, its mass con- [11] Chai S-P, Zein SHS, Mohamed AR. The effect of catalyst
tinues to increase. This process has not been reported previ- calcination temperature on the diameter of carbon
ously, and differs from the catalyst poisoning found in thin nanotubes synthesized by the decomposition of methane.
Carbon 2007;45(7):1535–41.
film catalyst growth. These results suggest that synthesis
[12] Endo M, Takeuchi K, Kobori K, Takahashi K, Kroto HW, Sarkar
durations for FC-CVD growth can be optimized to maximize
A. Pyrolytic carbon nanotubes from vapor-grown carbon
nanotube mat height and yield. fibers. Carbon 1995;33(7):873–81.
2094 CARBON 4 7 ( 2 0 0 9 ) 2 0 8 5 –2 0 9 4

[13] Serp P, Figueiredo JL. A microstructural investigation of [29] Sinnott SB, Andrews R, Qian D, Rao AM, Mao Z, Dickey EC,
vapor-grown carbon fibers. Carbon 1996;34(11):1452–4. et al. Model of carbon nanotube growth through chemical
[14] Dai H. Nanotube growth and characterization. In: vapor deposition. Chem Phys Lett 1999;315(1):25–30.
Dresselhaus MS, Dresselhaus G, Avouris P, editors. Carbon [30] Dresselhaus MS, Dresselhaus G, Jorio A, Souza Filho AG, Saito
nanotubes: synthesis, structure, properties, and applications, R. Raman spectroscopy on isolated single wall carbon
vol. 80. New York: Springer-Verlag. p. 29–53. nanotubes. Carbon 2002;40(12):2043–61.
[15] Dresselhaus MS, Endo M. Relation of carbon nanotubes to [31] Bacsa WS, Ugarte D, Chatelain A, de Heer WA. High-
other carbon materials. In: Dresselhaus MS, Dresselhaus G, resolution electron microscopy and inelastic light scattering
Avouris P, editors. Carbon nanotubes: synthesis, structure, of purified multishelled carbon nanotubes. Phys Rev B
properties, and applications, vol. 80. New York: Springer- 1994;50(20):15473–6.
Verlag. p. 11–28. [32] Tohji K, Goto T, Takahashi H, Shinoda Y, Shimizu N,
[16] Dresselhaus MS, Dresselhaus G, Sugihara K, Spain IL, Jeyadevan B, et al. Purifying single-walled nanotubes. Nature
Goldberg HA. Graphite fibers and filaments. New 1996;383(6602):679.
York: Springer-Verlag; 1988. p. 382. [33] Maultzsch J, Reich S, Thomsen C, Webster S, Czerw R, Carroll
[17] Yamada T, Namai T, Hata K, Futaba DN, Mizuno K, Fan J, et al. DL, et al. Raman characterization of boron-doped
Size-selective growth of double-walled carbon nanotube multiwalled carbon nanotubes. Appl Phys Lett
forests from engineered iron catalysts. Nat Nanotechnol 2002;81(14):2647–9.
2006;1:131–5. [34] Kelemen SR. The reaction of carbon and water as catalyzed
[18] Bai S, Li F, Yang Q-H, Cheng H-M, Bai J. Influence of ferrocene/ by a nickel surface. Appl Surf Sci 1987;28(4):439–74.
benzene mole ratio on the synthesis of carbon [35] McKee GSB, Vecchio KS. Thermogravimetric analysis of
nanostructures. Chem Phys Lett 2003;376(1–2):83–9. synthesis variation effects on CVD generated
[19] Sinnott SB, Andrews R, Qian D, Rao AM, Mao Z, Dickey EC, multiwalled carbon nanotubes. J Phys Chem B
et al. Model of carbon nanotube growth through chemical 2006;110(3):1179–86.
vapor deposition. Chem Phys Lett 1999;315:25–30. [36] Pan ZJ, Yang RT. Catalytic behavior of transition metal oxide
[20] Singh C, Shaffer MSP, Windle AH. Production of controlled in graphite gasification by oxygen, water, and carbon dioxide.
architectures of aligned carbon nanotubes by an injection J Catal 1991;130(1):161–72.
chemical vapour deposition method. Carbon [37] Thomas JM. Microscopic studies of graphite oxidation. In:
2003;41(2):359–68. Walker PL, editor. Chemistry and physics of carbon. New
[21] Lupo F, Rodrı́guez-Manzo JA, Zamudio A, Elı́as AL, Kim YA, York: Marcel Dekker, Inc.; 1966. p. 121–202.
Hayashi T, et al. Pyrolytic synthesis of long strands of large [38] Shimada T, Yanase H, Morishita K, Hayashi J-I, Chiba T.
diameter single-walled carbon nanotubes at atmospheric Points of onset of gasification in a multi-walled carbon
pressure in the absence of sulphur and hydrogen. Chem Phys nanotube having an imperfect structure. Carbon 2004;42(8–
Lett 2005;410:384–90. 9):1635–9.
[22] Aguilar-Elguézabal A, Antúnez W, Alonso G, Paraguay- [39] Wirth CT, Hofmann S, Robertson J, Mattevi C, Cepek C,
Delgado F, Espinosa F, Miki-Yoshida M. Study of carbon Goldoni A, et al. In situ time-resolved XPS study of catalyst
nanotubes synthesis by spray pyrolysis and model of growth. behavior during carbon nanotube growth. In: Proceedings of
Diam Relat Mater 2006;15:1329–35. the 2008 MRS spring meeting, San Francisco, CA, USA. 2008.
[23] Chaisitsak S, Nukeaw J, Tuantranont A. Parametric study of [40] Lodziana Z, Topsoe N-Y, Norskov JK. A negative surface
atmospheric-pressure single-walled carbon nanotubes energy for alumina. Nat Mater 2004;3(5):289–93.
growth by ferrocene–ethanol mist CVD. Diam Relat Mater [41] Shchipalov YK. Surface energy of crystalline and vitreous
2007;16:1958–66. silica. Glass Ceram (Trans Steklo i Keramika) 2000;57(11–
[24] Vander Wal RL, Ticich TM, Curtis VE. Substrate-support 12):374–7.
interactions in metal-catalyzed carbon nanofiber growth. [42] Liu H, Zhang Y, Arato D, Li R, Merel P, Sun X. Aligned multi-
Carbon 2001;39(15):2277–89. walled carbon nanotubes on different substrates by floating
[25] Bom D, Andrews R, Jacques D, Anthony J, Chen B, Meier MS, catalyst chemical vapor deposition: critical effects of buffer
et al. Thermogravimetric analysis of the oxidation of layer. Surf Coat Technol 2008;202(17):4114–20.
multiwalled carbon nanotubes: evidence for the role of defect [43] Sakata K, Honma K, Ogawa K, Watanabe O, Nii K. Interface
sites in carbon nanotube chemistry. Nano Lett chemistry and bonding strength for diffusion-bonded Fe/Fe–
2002;2(6):615–9. FeO/Al2O3 systems. J Mater Sci 1986;21(12):4463–7.
[26] Pang LSK, Saxby JD, Chatfield SP. Thermogravimetric analysis [44] Papirer E, Ligner G, Balard H, Vidal A, Mauss F. Surface energy
of carbon nanotubes and nanoparticles. J Phys Chem of silicas and c-alumina modified by heat treatment. Chem
1993;97(27):6941–2. Mod Surf 1990;3:15–26.
[27] Saxby JD, Chatfield SP, Palmisano AJ, Vassallo AM, Wilson MA, [45] Li Q, Zhang X, DePaula RF, Zheng L, Zhao Y, Stan L, et al.
Pang LSK. Thermogravimetric analysis of Sustained growth of ultralong carbon nanotube arrays for
buckminsterfullerene and related materials in air. J Phys fiber spinning. Adv Mater 2006;18:3160–3.
Chem 1992;96(1):17–8. [46] Puretzky AA, Geohegan DB, Jesse S, Ivanov IN, Eres G. In situ
[28] Bai S, Li F, Yang Q, Cheng H-M, Bai J. Influence of ferrocene/ measurements and modeling of carbon nanotube array
benzene mole ratio on the synthesis of carbon growth kinetics during chemical vapor deposition. Appl Phys
nanostructures. Chem Phys Lett 2003;376(1, 2):83–9. A 2005;81:223–40.

You might also like