You are on page 1of 8

ARTICLE IN PRESS YJCIS:14307

JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.1 (1-8)


Journal of Colloid and Interface Science ••• (••••) •••–•••

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science

www.elsevier.com/locate/jcis

Comparative study of carbon nanotube dispersion using surfactants


Richa Rastogi a,∗ , Rahul Kaushal a , S.K. Tripathi b , Amit L. Sharma a , Inderpreet Kaur a , Lalit M. Bharadwaj a
a
Biomolecular Electronics and Nanotechnology Division (BEND), Central Scientific Instruments Organisation (CSIO), Sector 30C, Chandigarh, India
b
Centre of Advanced Studies in Physics, Punjab University, Sector 14, Chandigarh, India

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

Article history: Dispersion of carbon nanotubes (CNTs) is a challenging task for their utilization in nanoscale device ap-
Received 8 April 2008 plications. This account reports a comparative analysis on dispersion of multiwalled carbon nanotubes
Accepted 1 September 2008 (MWNTs) with four surfactants—Triton X-100, Tween 20, Tween 80, and sodium dodecyl sulfate (SDS).
Among the four surfactants, Triton X-100 and SDS provide maximum and minimum dispersion, respec-
Keywords:
tively. Dispersion of MWNTs has been characterized with UV–vis spectroscopy and transmission electron
Dispersion microscopy (TEM). TEM results are in agreement with the UV–vis measurements. The experimentally ob-
Absorbance served trend of dispersing power of surfactants is consistent with their chemical structures. An optimum
Multiwalled nanotubes CNT-to-surfactant ratio has been determined for each surfactant. This parameter is shown to affect the
Surfactants nanotube dispersion significantly. Surfactant concentration above or below this ratio is shown to deteri-
orate the quality of nanotube dispersion. TEM analysis of a high-surfactant-concentration sample enables
us to construct a plausible mechanism for decrease in CNT dispersion at high surfactant concentration,
consistent with the UV–vis observations. Temperature stability of the surfactant is another important
factor affecting the quality of CNT dispersion.
© 2008 Elsevier Inc. All rights reserved.

1. Introduction aggressive chemical functionalization at high temperature creates


defects at the nanotube surface, consequently altering the electri-
Due to their tremendous optical, electrical, and mechanical cal properties of carbon nanotubes [9]. In contrast, a noncovalent
properties, carbon nanotubes enjoy a preeminent status in the approach involves adsorption of the chemical moieties onto the
panoply of nanomaterials, finding wide range of applications in nanotube surface, either via π –π stacking interaction such as in
biosensors [1], composites [2], field emission devices [3], elec- DNA, uncharged surfactants, etc., or through coulomb attraction in
tronic components [4], probe tips [5], etc. Delocalization of π - the case of charged chemical moieties. The noncovalent approach
electrons renders them conducting and alleviates adsorption of is superior in the sense that it does not alter the π -electron cloud
various chemical moieties via π –π stacking interaction [6]. A high of graphene, in turn preserving the electrical properties of carbon
aspect ratio makes them prone to entanglement and bundling. Par- nanotubes.
ticularly, carbon nanotubes are bundled with strong van der Waals Surfactants and polymers are extensively used for carbon nan-
interaction energy of ca. 500 eV/μm of tube–tube contact [7]. Such otube dispersion via noncovalent approach. Both of them get ad-
high interaction energy renders CNT dispersion a challenging task. sorbed onto nanotube surface, rendering them soluble in aqueous
Currently two approaches are widely used in nanotube disper- and organic solvents. Dispersion of nanotubes in polymer matri-
sion—the mechanical approach and the chemical approach. The ces may not be a proper choice for electronic device applications,
mechanical approach includes ultrasonication and high-shear mix- as polymer itself can participate in electrical events [10]. Disper-
ing. These processes are time-consuming and less efficient. Fur- sion using surfactants diminishes such anomalies, as they can be
thermore, as reported earlier by Lu et al. [8], ultrasonication can removed easily by washing. To date, a wide variety of surfactants
result in fragmentation of CNTs, in turn, decreasing their aspect ra- have been investigated for dispersion of carbon nanotubes, such
tio. Besides this, the stability of the dispersion is poor. On the other as sodium dodecyl benzenesulfonate (SDBS) [11], dodecyltrimethyl-
hand, the chemical approach includes both covalent and noncova- ammonium bromide (DTAB) [12], hexadecyltrimethylammonium
lent methods. Covalent methods involve functionalization with var- bromide (CTAB) [13], octyl phenol ethoxylate (Triton X-100) [14],
ious chemical moieties to improve solubility in solvents. However, and sodium dodecyl sulfate (SDS) [15]. In view of the large number
of surfactants available for dispersion, it is imperative to conduct
a systematic study of different parameters such as concentration,
* Corresponding author. Fax: +91 172 2657267. nature, stability, etc. in order to choose the right surfactant for a
E-mail address: richa.bend@gmail.com (R. Rastogi). particular application. Until recently, few attempts have been made

0021-9797/$ – see front matter © 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2008.09.015

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.2 (1-8)
2 R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–•••

to compare dispersing power of different surfactants. Moore et al. 2.3.2. Transmission electron microscopy
[16] have done comparative analysis of various surfactants on the The dispersions of MWNTs were also characterized using a
basis of their spectral properties. Hertel et al. [17] and Yurekli et al. transmission electron microscope (Hitachi H-7500) operating at
[18] have reported changes in phase behavior of carbon nanotubes 80 kV. MWNTs dispersed in surfactants were filtered, dried, and
on the basis of the nature, concentration, and type of interaction redispersed in distilled water at a concentration of 0.01 mg/ml.
of surfactants. Still, there is a lack of systematic study to optimize A drop of 15 μl of the above suspension was dropped onto carbon-
different parameters influencing dispersion of nanotubes. coated TEM grids (300 mesh, 3 mm, purchased from TAAB Labora-
The present study confronts a comparative analysis of four tories, England) and viewed under a transmission electron micro-
surfactants—Triton X-100, SDS, Tween 20, and Tween 80 [19–22]— scope.
for nanotube dispersion. The dispersing power of the surfactants
has been analyzed experimentally as well as theoretically on the 3. Results and discussion
basis of their structural organization. This study provides insights
into some parameters for optimization of dispersion of MWNTs 3.1. Comparison of dispersing power of surfactants using UV–vis
using surfactants. UV–vis spectroscopy and transmission electron spectroscopy
microscopy (TEM) have been employed to analyze the dispersion
ability of these surfactants. The significance of using a particular Bundled carbon nanotubes are not active in the UV–vis region
ratio of surfactants and MWNTs has been established for obtain- [23,24]. Only individual carbon nanotubes absorb in this region.
ing optimum dispersion, which may be cited as a relatively new Therefore, dispersion of carbon nanotubes can be characterized us-
finding in this area of research. From our study, the optimum CNT- ing UV–vis absorption spectroscopy. To characterize the dispersion
to-surfactant ratio turns out to be the most important parameter of MWNTs in surfactants using UV–vis spectroscopy, absorbance
in nanotube dispersion. We have optimized this ratio for these sur- values were recorded at 500 nm as reported in previous studies
factants. Apart from the above, temperature stability of surfactant [25–29]. This wavelength is virtually unaffected by ambient con-
is also shown to influence nanotube dispersion. ditions of nanotubes. The Lambert–Beer law is well obeyed by
MWNTs at this wavelength. Concentration of MWNTs dissolved or
2. Materials and method dispersed into the solution can then be determined using the spe-
cific extinction coefficient of carbon nanotubes at 500 nm, ε500 =
28.6 cm2 mg−1 [29], in the Lambert–Beer law. With this knowl-
2.1. Materials
edge, percentage recovery into the solution can be calculated as
c1
Purified MWNTs (diameter 40–70 nm, length 2–5 μm, purity % extractability = × 100, (1)
>95%) were procured from Amorphous & Nanostructure Ltd., USA. c
The surfactants—Triton X-100, Tween 20 (USB Corp., USA), Tween where c 1 = concentration of MWNTs recovered in solution and
80, and sodium dodecyl sulfate (Bio Basic, USA)—were used as re- c = concentration of MWNTs originally taken in surfactant. This
ceived. All solutions were prepared in 18 M cm deionized water. parameter is the measure of dispersion of carbon nanotubes in so-
lution.
2.2. Preparation of dispersion of MWNTs in various surfactants Fig. 1 depicts the UV–vis spectra of MWNTs with varying con-
centrations of nanotubes in surfactant solutions. The inset shows
the Lambert–Beer dependence of absorption at 500 nm on the
In order to compare the dispersing power of the four sur-
concentration of surfactant. Background absorption in spectra is
factants, dispersions of MWNTs were prepared at concentrations
due to the presence of bundles and mats of CNTs [13]. In order
spanning from 15 to 50 mg/L in steps of 5 mg/L, keeping the
to compare the dispersing power of surfactants in context, per-
concentration of surfactant (1%) constant. These 32 samples were
centage extractability was calculated at different concentrations of
ultrasonicated for 2 h in order to get surfactant-coated MWNTs.
nanotubes (Table S1 of the supplementary material).
Dispersions were analyzed with UV–vis spectroscopy and the max-
In all surfactants, percentage extractability vs concentration fol-
imum extractable concentration of MWNTs (at 1% surfactant con-
lows a Gaussian trend (Fig. 2), which depicts a linear increase
centration) was determined for each of the surfactants.
with increase in concentration of MWNTs until the maximum ex-
In order to find the optimum CNT-to-surfactant ratio for each
tractability limit is achieved. This is presumably because at low
surfactant, a second set of experiments were carried out. In these
CNT concentrations, the amount of surfactant is sufficient to coat
experiments, concentration of surfactants was varied from 1.1 to
the carbon nanotube surface evenly. Eventually, a concentration
1.9% in steps of 0.1%, keeping the amount of MWNTs constant.
value is attained for which the surfactant amount is just sufficient
Constant MWNT concentrations chosen in these experiments were
to disperse the carbon nanotubes; i.e., the maximum extractabil-
the maximum extractable concentrations of MWNTs (determined
ity limit is achieved at this point. For subsequent increases in
in the first set of experiments for 1% surfactant concentration).
the concentration of nanotubes, the surfactant amount turns in-
Again, these samples were analyzed using UV–vis spectroscopy.
sufficient to fully disperse the agglomerates of CNTs, therefore de-
creasing the percentage extractability at high concentrations. The
2.3. Characterization of dispersion maximum amount of MWNTs is extracted in the case of Triton X-
100, where percentage extractability has gone as high as 90.03%.
2.3.1. UV–vis spectroscopy It is estimated to be 89.98, 84.61, and 84.89% for Tween 80, SDS,
The dispersions of MWNTs in surfactants were characterized us- and Tween 20, respectively. It is noteworthy that maximum ex-
ing UV–vis spectrophotometer (Hitachi U-2800) operating between tractability limit is attained earlier for SDS in comparison to other
the ranges of 300–1100 nm. In the first set of experiments, baseline surfactants. Maximum extractable MWNT concentration is found
correction was carried out using pure 1% solutions of the four sur- to be 25, 30, 30, and 40 mg/L for SDS, Tween 20, Tween 80, and
factants so that their absorbance values got subtracted from that Triton X-100, respectively, for the constant surfactant concentra-
of MWNTs dispersions. In the second set of experiments, the base- tion (1%). Thus, the same amount of Triton X-100 can disperse
line was equilibrated every time a new sample was analyzed with large amounts of MWNTs as compared to other surfactants. Al-
corresponding surfactant concentration. though Tween 20 and Tween 80 show equivalent behavior in terms

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.3 (1-8)
R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–••• 3

Fig. 1. UV–vis spectra of carbon nanotubes in (a) SDS, (b) Tween 80, (c) Tween 20, and (d) Triton X-100 (Beer–Lambert curves inset).

of maximum extractable MWNTs concentration, Tween 80 proves adsorption and consequently dispersing power of surfactants are
to be better than Tween 20 in terms of percentage extractability. greatly affected by the tail length of the surfactant. Longer tails
Hence, according to our experimental results, the dispersing power means high spatial volume and more steric hindrance, thus provid-
of the four surfactants follows the trend ing greater repulsive forces between individual carbon nanotubes
[32]. Besides this, surfactants with unsaturated bonds in their tail
SDS < Tween 20 < Tween 80 < Triton X-100 groups contribute more toward nanotube dispersion [32].
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−→
Dispersing power As one can see, Tween 80 has the greatest hydrocarbon tail
length, while Triton X-100 has the smallest one (Fig. S1 of the
This experimentally observed trend for dispersing power of surfac- supplementary material). The phenyl ring has an effective length
tants can also be explained on the basis of their chemical struc- of about three and one-half carbon atoms, while carbon atoms
tures (Fig. S1 of the supplementary material). In order to disperse on branches contribute about one-half the effect of carbon atoms
nanotubes in water, surfactant molecules orient themselves in such on straight alkyl chains. Thus, Triton X-100 has an effective chain
a fashion that hydrophobic tail groups face toward the nanotube length of nine atoms only, which is the shortest out of the four.
surface while hydrophilic head groups face toward the aqueous Thus, theoretically Triton X-100 should exhibit minimum dispers-
phase, producing a lowering of the nanotube/water interfacial ten- ing power, contrary to experimental observations. Such a paradigm
sion. Thus, the dispersing power of the surfactant depends on departure from experimental observations is presumably because
how firmly it adsorbs onto the nanotube surface and produces of the presence of benzene ring in the tail group of Triton X-100.
by this adsorption energy barriers of sufficient height to aggre- This finding prompts us to conclude that whenever “tail length
gation. Molecules having the benzene ring structure adsorb more factor” and “benzene ring factor” compete, the latter contributes
strongly to the graphitic surface due to π –π stacking type interac- more to dispersion of CNTs. In total, Triton X-100 proves to be
tion [11,30]. Generally, hydrophobic tail groups tend to lie flat on the best among the four surfactants. Among the remaining three
the graphitic surface because graphitic unit cells match well with surfactants, Tween 80 has the longest tail. In addition, it has one
the methylene units of hydrocarbon chains [31]. Thus, efficiency of C–C unsaturated bond in its tail. Thus, theoretically, it is a better

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.4 (1-8)
4 R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–•••

Fig. 1. (continued)

dispersant than Tween 20 and SDS. Indeed, this is what has been trations were selected to be 25, 30, 30, and 40 mg/L for SDS,
observed experimentally. Among the remaining two, the tail of SDS Tween 20, Tween 80, and Triton X-100, respectively (chosen from
has an extra carbon atom as compared to Tween 20, but the head the first set of experiments). Again, absorbance values were de-
group of Tween 20 is bulkier than that of SDS. Bulkier head groups termined at 500 nm and percentage extractability of MWNTs was
cause larger steric stabilization, in turn, producing steric energy calculated for each sample (Table S2 of the supplementary mate-
barriers to aggregation in nonionic surfactants such as Tween 20 rial). There appears to be a polynomial correlation between per-
[33]. It is to be observed that these two factors are competing al- centage extractability and the concentration of surfactant (Fig. 3).
most equally when compared with the experimental results. There Percentage extractability attains maxima at 1.3 (Triton X-100), 1.2
is just a slight difference between maximum extractability of SDS (Tween 80), 1.4 (Tween 20), and 1.3% (SDS) surfactant concentra-
and Tween 20. This is presumably due to compensation for the tions. With this knowledge, optimum CNT-to-surfactant ratio was
effects of extra carbon atom in SDS by bulkier head group in calculated to be 1:466.66, 1:520, 1:400, and 1:350 for Tween 20,
Tween 20. Overall, therefore, structural considerations also support SDS, Tween 80, and Triton X-100, respectively. Such a large amount
our experimentally established trend of dispersing power among of surfactant required with respect to a small amount of MWNTs
these surfactants. can be attributed to the large surface-to-weight ratio of carbon
nanotubes.
3.2. Determination of optimum CNT to surfactant ratio using UV–vis Increase in percentage extractability with increase in surfactant
spectroscopy concentration was on the expected lines. However, the reason be-
hind the decrease in percentage extractability after attaining max-
In order to determine the optimum CNT-to-surfactant ratio ima was not clear. This problem was addressed by TEM observation
for each surfactant, a second set of experiments was carried out of a high-concentration surfactant sample (Fig. 4). Interestingly,
in which the concentration of surfactants was varied while the flocculation of CNTs was observed at some places. Individual CNTs
MWNT concentration was kept constant. Constant MWNT concen- were coated with thick nonuniform surfactant coatings. Chunks of

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.5 (1-8)
R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–••• 5

Fig. 2. Percentage extractability vs concentration trend of carbon nanotubes for (a) Tween 20, (b) SDS, (c) Tween 80, and (d) Triton X-100.

CNTs were aggregated via surfactant molecules. This observation ratio in different surfactants. In SDS dispersion, ribbon-like ropes of
can be explained in the light of the theory of micelle formation MWNTs were observed with bundle diameters ranging from 62.5
in surfactants. Fig. 5 depicts a schematic illustration of a plausi- to 218.75 nm. Organization of these ribbon-shaped ropes could not
ble mechanism of flocculation of CNTs via surfactant molecules. At be observed due to poor contrast. Small particles of diameter rang-
high concentrations, the surfactant molecules form micelles in so- ing between 70 and 100 nm were also observed. These are sug-
lution. The size of these micelles keeps on increasing with increas- gested to be SDS crystals and were present due to crystallization
ing surfactant concentration due to interaction between groups of SDS at low ambient temperature (17 ◦ C). Poor temperature sta-
of the same polarity. Likewise, surfactant molecules form multi- bility of SDS solution might also be the reason for poor dispersion
layers on nanotube surface when the concentration of surfactant of MWNTs in SDS. No entanglement of MWNTs was observed in
is increased for a constant nanotube concentration. As a conse- the cases of Tween 20, Triton X-100, and Tween 80. Well-isolated
quence, surface coverage by surfactant molecules becomes so high MWNTs were observed in these surfactants. Bundle diameters of
that portions of surfactant molecules extending into the liquid MWNTs ranged from ∼19.52 to 28.20 nm for Tween 20, ∼5.4
phase start interacting with others on neighboring CNTs [34]. If to 11.43 nm for Tween 80, and ∼4 to 5.35 nm for Triton X-100,
the orientation of the outermost layer is such that the hydropho- respectively. This finding reveals the highest dispersing power in
bic groups of surfactant molecules are forced to extend into the Triton X-100 and the minimum in SDS, consistent with UV–vis
aqueous phase, then this interaction favors a reduction in their observations. Thickness of surfactant coating was observed to be
surface energies [35]. This bridging of CNTs via extra surfactant ∼6.4–8 nm for Tween 20, ∼3–4 nm for Tween 80, and ∼0.9–1.78
molecules causes flocculation. Flocculation of CNTs might be the nm for Triton X-100, respectively. Stability of Tween 20, Tween 80,
reason behind the decrease in percentage extractability, which in and Triton X-100 at low ambient temperature might also be the
turn decreases the dispersion of nanotubes at high surfactant con- reason for better dispersion in these surfactants in comparison to
centration. SDS.

3.3. Morphology of CNT dispersions 4. Summary and conclusions

In order to characterize dispersion ability of different sur- For a comparative study of dispersion parameters of MWNTs,
factants morphologically, bundle diameters were measured using four surfactants, namely Triton X-100, Tween 20, Tween 80, and
TEM. Larger bundle diameter implies less dispersion. Fig. 6 de- SDS, were analyzed. The key motif of this study was to analyze
picts TEM micrographs of MWNTs at optimum CNT-to-surfactant the parameters responsible for the dispersion ability of a surfac-

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.6 (1-8)
6 R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–•••

Fig. 3. Variation of percentage extractability with variation of concentration of surfactant for (a) SDS, (b) Triton X-100, (c) Tween 80, and (d) Tween 20.

Fig. 5. Mechanism of flocculation of CNTs via surfactant molecules.

tant. With this objective in mind, UV–vis and TEM studies of the
dispersion of MWNTs in the above surfactants were executed. Dis-
persion of MWNTs in the four investigated surfactants shows the
following trend:
SDS < Tween 20 < Tween 80 < Triton X-100
−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−→
Dispersing power

The experimental trend was further correlated with the chem-


ical structures of the surfactants. Our experimental investigation
shows unambiguously that, contrary to the theoretical aspect, Tri-
ton X-100 enjoys the highest dispersing power among four sur-
Fig. 4. Transmission electron micrograph of a high-surfactant-concentration sample. factants by virtue of its benzene ring. This analysis enables us

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.7 (1-8)
R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–••• 7

Fig. 6. Transmission electron micrographs of carbon nanotube dispersion in (a) SDS, (b) Tween 80, (c) Tween 20, and (d) Triton X-100.

to conclude that the “benzene ring factor” enjoys reigning status assistance. The authors are also thankful to the Director, Central
in comparison to the “tail length factor” in dispersion of MWNTs Scientific Instruments Organization (CSIO), Chandigarh, for provid-
using surfactants. Therefore, the dispersing power of different sur- ing necessary infrastructure and support and to Mr. M.L. Sharma
factants is inherent, due to their chemical organization. The key of the Regional Sophisticated Instrumentation Center (RSIC), Panjab
finding of the present study is the significance of optimum CNT- University, Chandigarh for transmission electron microscope facil-
to-surfactant ratio. The quality of nanotube dispersion deteriorates ity.
below or above this ratio. Thus, surfactants should be in concentra-
tion just sufficient to coat the nanotube surface, avoiding any ex- Supplementary material
cess, as an unnecessarily large amount of surfactant also decreases
the nanotube dispersion. To ascertain the reason for this decrease The online version of this article contains additional supple-
in the dispersion, a TEM study was employed. The TEM study mentary material.
permits the elucidation of a plausible mechanism indicating de- Please visit DOI: 10.1016/j.jcis.2008.09.015.
creasing dispersion at high surfactant loading to be caused by the
flocculation of CNTs with undesired excess surfactant molecules. References
Furthermore, a low degree of dispersion of CNTs at low concentra-
tions of surfactant is attributed to an insufficient reagent amount. [1] X. Tang, S. Bansaruntip, N. Nakayama, Y. Erhan, Y.L. Chang, Q. Wang, Nano
We have standardized this ratio for the surfactants in context. It Lett. 6 (2006) 1632.
was found to be 1:466.66, 1:520, 1:400, and 1:350 for Tween 20, [2] L. Clayton, T. Gerasimov, M. Meyyappan, J.P. Harmon, Adv. Funct. Mater. 15
(2005) 101.
SDS, Tween 80, and Triton X-100, respectively. Nanotube dispersion [3] J. Koohsorkhi, Y. Abdi, S. Mohajerzadeh, H. Hosseinzadegan, Y. Komijani, E.A.
was found to be highest at these ratios. Moreover, thermal stabil- Soleimani, Carbon 44 (2006) 2797.
ity of surfactant solution also plays an important role in dispersion [4] E. Frackowiak, F. Beguin, Carbon 40 (2002) 1775.
of MWNTs. Surfactants such as SDS disperse the nanotubes poorly [5] H.J. Dai, J.H. Hafner, A.G. Rinzler, D.T. Colbert, R.E. Smalley, Nature 384 (1996)
147.
due to instability of the solution at low temperature. In total, one
[6] M. Terrones, Annu. Rev. Mater. Res. 33 (2003) 419.
can conclude that proper choice of a suitable surfactant needs con- [7] L.A. Girifalco, M. Hodak, R.S. Lee, Phys. Rev. B 62 (2000) 13104.
sideration of its structure, its optimum ratio to nanotubes, and [8] K.L. Lu, R.M. Lago, Y.K. Chen, M.L.H. Green, P.J.F. Harris, S.C. Tsang, Carbon 34
thermal stability of its solution. (1996) 814.
[9] J. Hilding, E.A. Grulke, Z.G. Zhang, F. Lockwood, J. Dispers. Sci. Technol. 24
(2003) 1.
Acknowledgments
[10] T. Blythe, D. Bloor, Electrical Properties of Polymers, Cambridge University
Press, London, 2005.
The authors thank the Department of Information and Technol- [11] M.F. Islam, E. Rojas, D.M. Bergey, A.T. Johnson, A.G. Yodh, Nano Lett. 3 (2003)
ogy (DIT), Ministry of Science and Technology, India, for financial 269.

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015
ARTICLE IN PRESS YJCIS:14307
JID:YJCIS AID:14307 /FLA [m5G; v 1.65; Prn:9/10/2008; 11:33] P.8 (1-8)
8 R. Rastogi et al. / Journal of Colloid and Interface Science ••• (••••) •••–•••

[12] E.A. Whitsitt, A.R. Barron, Nano Lett. 3 (2003) 775. [25] W. Huang, S. Taylor, K. Fu, Y. Lin, D. Zhang, T.W. Hanks, A.M. Rao, Y.P. Sun, Nano
[13] A.G. Ryabenko, T.V. Dorofeeva, G.I. Zvereva, Carbon 42 (2004) 1523. Lett. 2 (2002) 311.
[14] H. Wang, W. Zhou, D.L. Ho, K.I. Winey, J.E. Fischer, C.J. Glinka, E.K. Hobbie, Nano [26] K.D. Ausman, R. Piner, O. Lourie, R.S. Ruoff, M. Korobov, J. Phys. Chem. B 104
Lett. 4 (2004) 1789. (2000) 8911.
[15] J. Yu, N. Grossiord, C.E. Koning, J. Loos, Carbon 45 (2007) 618. [27] O.K. Kim, J. Je, J.W. Baldwin, S. Kooi, P.E. Pehrsson, L.J. Buckley, J. Am. Chem.
[16] V.C. Moore, M.S. Strano, E.H. Haroz, R.H. Hauge, R.E. Smalley, Nano Lett. 3
Soc. 125 (2003) 4426.
(2003) 1379.
[28] V.A. Sinani, M.K. Gheith, A.A. Yaroslavov, A.A. Rakhnyanskaya, K. Sun, A.A.
[17] T. Hertel, A. Hagen, V. Talalaev, K. Arnold, F. Hennrich, M. Kappes, Nano Lett. 5
Mamedov, J.P. Wicksted, N.A. Kotov, J. Am. Chem. Soc. 127 (2005) 3463.
(2005) 511.
[18] K. Yurekli, C.A. Mitchell, R. Krishnamootri, J. Am. Chem. Soc. 126 (2004) 9902. [29] A. Ikeda, T. Hamano, K. Hayashi, J. Kikuchi, Org. Lett. 8 (2006) 1153.
[19] J. Liu, A.G. Rinzler, H.J. Dai, J.H. Hafner, R.K. Bradley, P.J. Boul, A. Lu, T. Iver- [30] J.F. Liu, W.A. Ducker, Langmuir 16 (2000) 3467.
son, K. Shelimov, C.B. Huffman, F.J. Rodriguez-Marcias, Y.S. Shon, T.R. Lee, D.T. [31] D.M. Cyr, B. Venkataraman, G.W. Flynn, Chem. Mater. 8 (1996) 1600.
Colbert, R.E. Smalley, Science 280 (1998) 1253. [32] D.H. Napper, Polymeric Stabilization of Colloidal Dispersion, Academic Press,
[20] M.J. O’Connell, S.M. Bachilo, C.B. Huffman, V.C. Moore, M.S. Strano, E.H. Haroz, London, 1983.
Science 297 (2002) 593. [33] V.C. Moore, M.S. Strano, E.H. Haroz, R.H. Hauge, R.E. Smalley, Nano Lett. 3
[21] L. Jiang, L. Gao, J. Sun, J. Colloid Interface Sci. 260 (2003) 89. (2003) 1379.
[22] W. Wenseleers, I.I. Vlasov, E. Goovaerts, E.D. Obraztsova, A.S. Lobach, A.
[34] M.J. Rosen, Surfactants and Interfacial Phenomena, Wiley–Interscience, New
Bouwen, Adv. Funct. Mater. 14 (2004) 1105.
York, 1978.
[23] Y. Junrong, N. Grossiord, C.E. Koning, J. Loos, Carbon 45 (2007) 618.
[24] J.S. Laurent, C. Voisin, G. Cassabois, C. Delalande, P. Roussignol, O. Jost, L. Capes, [35] P. Somasundaran, T.W. Healy, D.W. Fuerstenau, J. Colloid Interface Sci. 22 (1966)
Phys. Rev. Lett. 90 (2003) 057404. 599.

Please cite this article in press as: R. Rastogi et al., J. Colloid Interface Sci. (2008), doi:10.1016/j.jcis.2008.09.015

You might also like