Вы находитесь на странице: 1из 10

Article

pubs.acs.org/Macromolecules

Viscoelasticity of Single-Walled Carbon Nanotubes in Unsaturated


Polyester Resin: Eects of Purity and Chirality Distribution
Joyanta Goswami and Virginia A. Davis*
Department of Chemical Engineering, Auburn University, Auburn, Alabama 36849, United States
S Supporting Information
*

ABSTRACT: The recent commercialization of single-walled carbon nanotubes


(SWNT) with controlled chirality distributions has created new opportunities for
producing SWNT materials with tailored electrical properties. However, there has
been relatively little research on understanding the eects of chirality distribution
and SWNT purity on the two main determinants of nanocomposite mechanical
properties: dispersion microstructure and nanotuberesin interactions. To
establish a framework for comparing dispersion and interactions, we investigated
the viscoelasticity of four SouthWest NanoTechnologies SWNT products: a low
and high purity semiconducting grade and a low and high purity metallic grade.
Optical microscopy of the dispersions did not show any signicant dierences
between the SWNT types. However, analysis of the dispersions viscoelastic
properties revealed the dierence in dispersion microstructure and the relative
strength of SWNTresin interactions. While all four products had a similar percolation threshold, the concentration of nonSWNT carbon impurities had a greater eect than chirality on viscoelastic properties and the relative strength of resinSWNT
interactions.

INTRODUCTION
Polymer nanocomposite properties are a function of intrinsic
material properties, dispersion microstructure, and nanomaterialpolymer interaction. For carbon nanotube (CNT) composites, achieving homogeneous dispersion of individual or even
small bundles of nanotubes can be a signicant challenge; the
attractive interactions are on the order of 104 kBT for singlewalled carbon nanotubes (SWNT).1,2 Rheology is a powerful
probe of both dispersion state and nanomaterialpolymer
interactions since it provides quantitative information on bulk
samples and can be used regardless of optical opacity or size
distribution.310 The majority of work on understanding CNT
dispersion has focused on understanding the impact of dramatic
structural dierences (i.e., single versus multiwalled carbon
nanotubes), intentional surface chemistry modications, or
mixing methodologies (e.g., sonication, incipient wetting, and
high shear). In the past few years, there has been signicant
progress in the chirality selective synthesis and postsynthesis
sortation of SWNT. For example, SouthWest NanoTechnologies Inc. sells several distinct SWNT products with dierent
chirality distributions. Chirality refers to the (n, m) vector that
denotes the orientation of the rolled graphene sheet comprising
the SWNT. Chirality is a well-known determinant of diameter,
electrical, and optical properties. For example, when n = m, the
nanotubes are known as armchair SWNT and have ballistic
electrical conduction. When m = 0, the SWNT are referred to
as zigzag, and when m n 0, the nanotubes are referred to as
chiral. When n m is divisible by 3, the SWNT are metallic or
low band gap semiconductors, and when n m is not divisible
by 3, the SWNT are semiconductors. Historically, most SWNT
2015 American Chemical Society

synthesis schemes produced a wide chirality distribution.


Advances in chirality controlled synthesis, and post-synthesis
chirality sortation, have largely been driven by the potential for
controlled electrical and optical properties. The impact of
controlled chirality distributions on mechanical properties of
bulk materials has not been thoroughly investigated. The eect
of chirality on the mechanical properties of individual SWNT is
well established,12 but the actual impact on bulk materials such
as polymer nanocomposites is less understood. In general, bulk
nanocomposite mechanical properties are determined by the
degree of dispersion and SWNTpolymer interactions. These
two parameters are complex functions of numerous factors
including the length to diameter ratio, concentration, surface
chemistry of the nanotubes, and processing parameters.4,5,8,10,13
In the case of thermosets, dispersion microstructure can also
aect curing, which plays a signicant part in nal nanocomposite properties.
In this research, dispersions of four commercial SWNT
products in unsaturated polyester resin (UPR) were compared
in order to understand the eects of purity and chirality on
viscoelasticity. UPR was chosen as the resin based on our
previous research,10,14 its continued dominance of the global
thermoset market, and the fact that there has still been limited
research on improving UPR properties via the incorporation of
SWNT. The four SWNT products obtained from SouthWest
NanoTechnologies were CG200, CG300, SG65, and SG65i.
Received: May 2, 2015
Revised: August 17, 2015
Published: November 16, 2015
8641

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules
Table 1. Summary of Key SWNT Properties
SWNT grade

SWNT (wt %)

non-SWNT C (wt %)

catalyst (wt %)

CG200 Lot. 14
CG300 Lot. 32
SG65 Lot. 35
SG65i Lot. 48

85
95
69
94

8
2
23
3

7
2
8
3

Raman D/G 785, 514 nm


0.07,
0.05,
0.14,
0.11,

0.06
0.04
0.07
0.03

av diam (nm)

av length (nm)

1.00
0.96
0.95
0.82

470
427
496a
498

Length estimated by scaling the average length from SWNT-dsDNA supernatants by the average C12-SWNT:dsDNA-SWNT length ratio for
CG200 and SG65i.

Figure 1. Optical microscopy images of SWNT-UPR dispersions (40 magnications stitched over an area of 0.97 0.73 mm). Top row (a, b, c, d)
0.05 0.005 vol %. Bottom row (e, f, g, h) 0.23 0.02 vol %. Scale bar: 100 m.
were an accumulation of 10 runs with exposure time of 10 s for each
run. SWNT surface chemistry was characterized by X-ray photoelectron spectroscopy (XPS) using a Kratos Analytical XSAM 800.
Dispersion Preparation. All calculations were based on densities
of 1.14 g/cm3 for UPR, 1.55 g/cm3 for CG200 and CG300 (actual
SWNT), and 1.59 g/cm3 for SG65and SG65i (actual SWNT) in a total
volume of 28 mL. The dispersions were mixed in accordance with our
previously published protocols.10,14 The required amount of SWNT
was weighed into a 50 mL round bottomed ask, and then the
appropriate amount of resin was added. The dispersions were mixed in
an Ace Glass (Vineland, NJ) Trubore four-neck ask head with Teon
stirring shaft attached to a exible shaft. Before starting the shear
mixing, the ask was purged with argon, and all joints and opening
were sealed with Paralm. Mixing was conducted at approximately
1000 rpm for 3 days while the ask was kept partially submerged in a
cold water bath to minimize styrene evaporation. Dispersions were
stored in sealed vials after removal from the mixer. The mixtures were
prepared with the goal of achieving equivalent amounts of actual
SWNT in the CG200, CG300, SG65, and SG65i dispersions; the
sample masses were dierent due to dierences in purity and chirality
(density) distribution. All samples were within 0.03 vol % of the
target concentration. For CG200, the actual concentrations were
0.010, 0.020, 0.050, 0.090, 0.13, 0.17, and 0.22 vol %. For SG65 the
actual concentrations were 0.010, 0.020, 0.050, 0.100, 0.14, 0.19, and
0.25 vol %. The dispersions prepared with CG300 the concentrations
were 0.050 (CG300 Lot # 15), 0.100, 0.14, 0.17, and 0.20 vol % and
with SG65i were 0.05, 0.10, 0.14, 0.17, and 0.21 vol %.
Dispersion Characterization. Dispersions were characterized
24 h after removal from the mixer. Microscopy was performed using
a Nikon Eclipse 80i optical microscope (Melville, NY) using LU Plan
Fluor 20/0.45 NA Nikon objective, NIS Elements software, and a
Prior scanning stage to enable image stitching. Rheology was
performed using an Antor Paar (GmbH) Physica MCR 301 rheometer
at a temperature of 10 C. The temperature was controlled by a Peltier
temperature control device. A cone and plate measuring geometry with
diameter 50 mm and cone angle 2.018 was used with a gap setting of
0.052 mm between the center of upper and lower plates. In addition to
three sample loadings in cone and plate geometry for all
concentrations, a 50 mm parallel plate geometry was also used for
some SWNT concentrations to ensure the rheology data were

The chirality distribution of CG200 contains a relatively high


fraction of metallic SWNT; CG300, a more recently
commercialized analogue, has a narrower distribution and a
lower fraction of non-SWNT carbon impurities. In contrast,
SG65 has a relatively high fraction of non-SWNT carbon but
contains predominantly semiconducting SWNT chiralities,
particularly (6, 5). SG65i has a similar chirality distribution as
SG65 but the purity of CG300. At least ve dispersions of each
product in UPR were prepared at comparable volume fractions
of SWNT. Careful characterization of the dispersions
viscoelasticity provided information on microstructure, the
formation of a percolated network, the relative strength of the
nanomaterial resin interaction, and dispersion stiness.
Chirality distribution had little eect on the SWNTresin
viscoelasticity and microstructure. The results highlight how
detailed analysis of rheological parameters can provide insights
into subtle dierences between the dispersions of similar
materials.

EXPERIMENTAL SECTION

Raw Materials. Reichhold Polylite 31830-00 UPR (70.5 wt %


isophthalic polyester/29.5 wt % styrene) was purchased from
Plasticare, LLC (Englewood, CO). This Newtonian resin had a
viscosity of 3.1 Pa s at 10 C. SWNT products SG65 (Lot # 35),
CG200 (Lot # 14), SG65i (Lot # 48), and CG300 (Lot # 32 and # 15)
were obtained from SouthWest NanoTechnologies Inc. (Norman,
OK) and used as received. The SWNT compositions were measured
in air using a TA Instruments (New Castle, DE) Q50 TGA. The
temperature was ramped from room temperature to 120 C at 10 C/
min and then held for 20 min to remove moisture and then ramped at
the same rate to 800 C followed by a 45 min hold. The weight loss
was analyzed using the published Sigma-Aldrich method to determine
the percentage of actual SWNT, non-SWNT carbon, and catalyst.15 A
Pacic Nanotechnology Inc. (Santa Clara, CA) Nano-R AFM was used
to characterize length and diameter distributions for C12 functionalized SWNT.16,17 Raman spectra were obtained using both 514 and
785 nm lasers on a Renishaw inVia Raman microscope (Homan
Estates, IL) linked with a Leica 50 (0.75 NA) objective. The spectra
8642

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

Figure 2. Steady shear rheology curves of viscosity as function of shear rate for SWNT-UPR dispersions at dierent SWNT concentrations; viscosity
vs shear rate with Sisko model t (a, b, c, and d). Concentrations: 0.05 (0.005), 0.095 (0.005), 0.135 ( 0.005), 0.18 ( 0.01) and 0.23 ( 0.02)
vol %. Error bars: < 9 %.
independent of xture-related artifacts. For each dispersion, at least
two rheometer loadings were characterized. The rst consisted of an
amplitude sweep to the onset of nonlinear viscoelasticity c followed
by a frequency sweep at < c; this sequence was repeated and
followed by a steady shear transient tests to determine the time to
steady state. The second load repeated the amplitude and frequency
sweep sequence but was followed by a steady shear ow curve from
0.01 to 100 s1. Based on the transient tests, the sampling time for all
samples was logarithmically decreased with shear rate from 1500 s per
measurement at 0.01 s1 to 50 s at 100 s1. Error bars for all
measurements were a maximum of 12% even for measurements made
using dierent test sequences or repeated amplitude sweep testing
over 5 h.

aggregates were present. With increasing concentration, the


SWNT-rich phase became more dominant, more aggregates
were observed, and the SWNT network completely lled the
image at a concentration of = 0.23 0.02 vol %.
Steady Shear Rheology. While optical microscopy is
useful for probing dramatic dierences between dispersions,
rheology is a much more sensitive, quantitative method.
Generally, in the absence of lyotropic phase behavior, higher
viscosity is indicative of better dispersion. In addition, steady
shear viscosity has practical implications for transporting and
processing dispersions. Figure 2 shows that steady shear
viscosity measurements enable more distinctions between
samples than optical microscopy. All the dispersions exhibited
both a low shear viscosity that was signicantly higher than the
3.1 Pa s Newtonian viscosity of the UPR and shear thinning
behavior. CG200 displayed lower steady shear viscosities than
the other SWNT types at all concentrations except 0.05 vol %
where it was higher than SG65. Another notable dierence
between the dispersions was that the two highest CG200
concentrations exhibited some evidence of a low shear
Newtonian plateau. For a given concentration, steady shear
viscosities for the CG300 and SG65i dispersions were always
greater than those for their lower purity versions (CG200 and
SG65). Although it was less pronounced for CG200, all samples
showed a signicant viscosity increase from 0.05 to 0.10 vol %.
At the higher concentrations, CG300 had a markedly higher
viscosity than even SG65i. For example, at 0.22 0.02 vol %,
the low shear viscosities (0.01 s1) for CG200, CG300, SG65,
and SG65i were 1460, 5390, 2600, and 4450 Pa s, respectively.
The higher viscosities for the CG300 and SG65i relative to
their lower purity versions (CG200 and SG65) are attributed to
the higher purity grades having fewer, lower aspect ratio
impurities. It is interesting that the viscosity of CG300 is
greater than that for SG65i while the viscosity of CG200 is less
than that of SG65. This may simply be due to the large fraction

RESULTS AND DISCUSSION


Initial Comparison. Table 1 summarizes key properties of
CG200, CG300, SG65, and SG65i. XPS showed no signicant
dierences in SWNT surface chemistry. Both CG200 and SG65
had 8 wt % catalyst residue while CG300 and SG65i had 3
wt % catalyst residue; these values are comparable to previous
work on other types of puried SWNT. However, the SG65
had a much higher percentage of non-SWNT carbon (MWNT,
graphite, carbon bers)18 and a higher D/G ratio than the other
products. The TGA plots and Raman spectra are provided in
Figures S1 and S2 of the Supporting Information. The samples
also had similar size distributions. The large carbon impurities
in SG65 were a signicant hindrance to meaningful AFM
imaging of C12-SWNT. Measurements on aqueous dsDNA
stabilized SG65 supernatants yielded clearer images. Since
dsDNA dispersion requires tip sonication which shortens
SWNT, the length of the SG65 was estimated based on the
ratio of the measured C12-SWNT length to dsDNA-SWNT
length for the other samples. Optical microscopy of the
dispersions did not reveal any obvious dierences (Figure 1)
between the SWNT dispersion states at a given concentration.
At low concentration, gray SWNT-rich regions were
interspersed with lighter resin-rich regions; a few dense black
8643

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

SG65i had similar values of G and c. At all concentrations,


SG65 exhibited lower G and higher c than the other SWNT
types; this is attributed to the much higher quantity of nonSWNT carbon. The tan values provide further evidence of
SG65 having more uid behavior than the other materials; for
SG65 dominant solid behavior was observed at 0.14 vol %, and
for all other SWNT types it started at 0.10 vol %.
Frequency sweep tests at amplitudes below c further
claried dierences in the viscoelasticity of the SWNT
dispersions. Plots of G and G delineate the solid-like
contribution of the SWNT network and viscous nature of the
polymer matrix, respectively, in accordance with the two-phase
contribution to viscoelasticity described by Filippone et al.21
Figure 3 shows changes in G as a function of frequency and
concentration for the four SWNT-UPR systems; data for G
are provided in Figure S5. Even at 0.05 vol %, G and G
deviated from terminal behavior (G 2, G ).22 The low
frequency slopes for G and G, and the numerical values of G
at = 0.01 rad/s, are shown in Table 3. At 0.05 vol %, G ( =
0.01 rad/s) was 11 1 Pa for CG200, CG300, and SG65i;
these values were over 3 times higher than that of SG65. At
0.095 vol % and above, the values for CG200 were lower than
those for CG300 and SG65i but still signicantly higher than
those for SG65. There was a marked increase in G between
0.050 and 0.095 0.005 vol % for all four SWNT. The Cole
Cole plots in Figure 4 further highlight that these concentration
had markedly dierent microstructures, while there were less
pronounced microstructural dierences between higher concentrations.
As described by Kayatin and Davis14 and Chen et al.,23
master curves for viscoelastic properties were constructed by
scaling the moduli by the corresponding values at the crossover
point (Figure 5). Above c the properties are dominated by the
viscous UPR, and below c they are dominated by the SWNT
network. The striking collapse indicates that all viscoelastic
moduli have a similar functional form, and the magnitudes are a
function of cluster size increasing with concentration.23 Trappe
and Weitz24 established a similar scaling methodology based on
using the low frequency moduli instead of crossover values.
This method is best suited for concentrations above
percolation. As shown in Figure S6, this method gives less
striking collapse due to inaccuracies in measuring G as 0
for low concentrations.
Percolation and Network Characteristics. There are
several established methods for estimating the critical
concentration for rheological percolation c; the results from
several methods are typically compared to provide the most
accurate value for c. The simplest method is inspection of the
G versus frequency plots. Both an order of magnitude increase
and a plateau in G versus are often used to provide a rst
approximation for c; based on these criteria, inspection of
Figure 3 suggests that c 0.1 vol % for all four SWNT.
Pogodina and Winter25 and Romasanta et al.26 used plots of tan
versus to identify the formation of a percolated network.
With increasing concentration, G increases at a faster rate than
G, resulting in tan = G/G < 1. Upon network formation,
low frequency tan plateau emerges.25 Figure 6 shows a clear
low frequency (0.010.1 rad/s) tan plateau exists for CG200,
CG300, and SG65i at 0.1 vol %, but a plateau was less
evident for SG65. The plateau behavior suggests c is
approximately 0.1 vol % but slightly higher for SG65. A similar
result was obtained by estimating percolation based on the
divergence of the complex modulus G* versus complex

of non-SWNT carbon in SG65, resulting in a much larger total


carbon mass in the SG65 than in the CG200. At high shear,
ow alignment results in the viscosity approaching that of the
Newtonian UPR, resulting in curvature of the data. This
curvature can be removed by plotting the specic viscosity sp =
( s)/s (Figure S3); reduced viscosity plots are also shown
in the Supporting Information (Figure S4). The lines in Figure
2 represent the Sisko model t = + ky(n1). The Sisko model
simply adds the innite shear viscosity to the standard
power law model where k and n are the consistency and power
law indices, respectively. This model worked best for
intermediate concentrations and was chosen over the more
commonly used Cross or CarreauYasuda models due to the
lack of a well-dened plateau in all data. All model parameters
are given in Tables S1S4.
Viscoleasticity. Comparing viscoelastic behavior measured
by oscillatory shear rheology has proven to be a powerful tool
for understanding SWNT dispersion behavior.6,7,10,19,20 In
addition, viscoelasticity provides a qualitative indication of the
mechanical properties of the nal composites. Viscoelasticity
measurements can determine the rheological percolation
threshold, the concentration at which the nanotubes form a
polymer spanning network that can provide mechanical
property enhancement. Additional details about the dispersion
microstructure and relative strength of nanotubenanotube
and nanotubepolymer interactions can also be discerned
through analysis of viscoelastic properties. For example, Kayatin
and Davis as well as Chatterjee and Krishnamoorti showed
superposition of viscoelastic moduli due to the similar nature of
CNT structure growth with concentration.14,19 Viscoelastic
property relations were also used to calculate fractal characteristics of nanotube dispersions by Khalkhal et al. and UrenaBenavides et al.7,10 Filiponne and Luna extended these and
similar superposition approaches to propose a universal
approach for understanding viscoelasticity in polymer composites.21
In this research, the linear viscoelastic region for each
dispersion was determined by probing increasing amplitude at a
constant frequency of 10 rad/s; a 2% decrease in G was used to
dene the critical strain c. Table 2 displays the linear
viscoelastic behavior for the SWNT-UPR dispersions. Elastic
behavior increased with SWNT concentration, and the strain
limit for the linear regime decreased. At similar concentrations,
G values for CG200 were typically lower than for CG300 and
SG65i but appreciably higher than those for SG65. CG300 and
Table 2. Comparison of Linear Viscoelastic Regimes in UPR
CG200
vol %
0.045
0.092
0.13
0.17
0.22

c (%)

G (Pa)

1.66
55
1.16
177
0.68
400
0.66
475
0.44
1020
SG65

CG300
tan

vol %

c (%)

1.90
0.93
0.63
0.59
0.44

0.05
0.10
0.14
0.17
0.20

2.25
0.99
0.81
0.61
0.50

G (Pa)

67
283
475
868
1280
SG65i

tan
1.90
0.81
0.66
0.56
0.48

vol %

c (%)

G (Pa)

tan

vol %

c (%)

G (Pa)

tan

0.05
0.10
0.14
0.19
0.25

3.11
1.38
1.11
1.00
0.71

24
168
220
373
577

3.00
1.70
0.85
0.72
0.62

0.05
0.10
0.14
0.17
0.21

2.30
0.92
0.63
0.62
0.45

57
271
488
724
1150

2.00
0.82
0.66
0.59
0.47
8644

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

Figure 3. Oscillatory shear rheology of SWNT-UPR: variation of G with angular frequency at dierent SWNT concentrations. Concentrations: 0.05
(0.005), 0.095 (0.005), 0.135 ( 0.005), 0.18 ( 0.01) and 0.23 ( 0.02) vol %. (Error < 12 %).

nor the signicant dierences in the amount of non-SWNT


carbon impurities had a discernible eect on percolation.
In addition to being one of the most common methods for
determining c, the relation Go ( c) also provides
qualitative information about dispersion microstructure. The
magnitude of denes the networks response to stressinduced deformations. In simulation work on percolated elastic
networks, Arbabi and Sahimi27 categorized elastic networks that
undergo attraction induced occulation and found 2.10 is
indicative of particle bonds that are resistant to stretching but
free to rotate and 3.75 is indicative of networks where both
bond stretching and rotation were resisted. Surve et al.
extended this to polymer nanocomposites and found
1.88 when network formation is assisted by polymer bridging
and there are signicant polymernanoparticle interactions.28
Filippone et al. compared several nanoparticlepolymer
systems and concluded that higher values of result from a
higher degree of nanoparticlenanoparticle interactions in
comparison to nanoparticlepolymer interactions. In other
words, lower values of indicate better nanoparticlepolymer
interaction.21 Similar conclusions were proposed by Surve et al.
in their comparison of several polymerller systems.28 Based
on the range of values c for the SWNT-UPR systems in this
research, 1.15 < < 1.27; this compares to the previously
reported result of = 1.55 for puried HiPco SWNT-UPR.14
The puried HiPco SWNT have a very broad chirality and
diameter distribution but no appreciable non-SWNT carbon.
These results may suggest better nanomaterialpolymer
interaction in the case of the SouthWest NanoTechnology
SWNTs. For SWNTPMMA melts, Du et al. found an even
lower value of = 0.70 (0.92 when converted to volume
percent by Filippone and de Luna).4,21 Together these ndings
show that even though nanotubes are strongly attracted to each
other, a range of SWNTpolymer attractions can be achieved
even in the absence of intentional functionalization.

Table 3. Low Frequency Slopes of G and G and Values of


G at = 0.01 s1.
CG200

CG300

vol %

slope G

slope G

G
(Pa)

0.05
0.09
0.13
0.17
0.22

0.10
0.09
0.08
0.08
0.08

0.30
0.10
0.08
0.09
0.08

12
65
180
212
485

vol %

slope G

slope G

G
(Pa)

0.05
0.10
0.14
0.17
0.20

0.09
0.07
0.07
0.07
0.07

0.40
0.10
0.10
0.07
0.06

11
114
212
416
617

SG65

SG65i

vol %

slope G

slope G

G
(Pa)

vol %

slope G

slope G

G
(Pa)

0.05
0.01
0.14
0.19
0.25

0.10
0.09
0.07
0.07
0.06

0.50
0.30
0.20
0.10
0.08

3
45
86
161
270

0.05
0.10
0.14
0.17
0.21

0.08
0.06
0.06
0.07
0.06

0.40
0.10
0.07
0.06
0.06

10
110
234
355
597

viscosity * based on the method described by Mitchell et al.


and others.8,10,14 The x-intercepts of the linear regions of the
plots of G* vs (Figure 7ad) resulted in c values of 0.082
(R2 = 0.92), 0.086 (R2 = 0.95), 0.082 (R2 = 0.98), and 0.082 (R2
= 0.97) vol % for CG200, CG300, SG65, and SG65i,
respectively. Percolation was also estimated by the frequently
used method of tting the power law relation G0 ( c)
by adjusting c values in the percolation range to maximize the
R2 value (Figure 7e,f). This resulted in slightly lower estimates
of c: 0.060 vol % for CG200 and 0.065 vol % for the other
SWNT types. Comparison of the results shows that 0.06 c
0.10. This result is consistent with c R/L, which suggests
many SWNT are dispersed as individuals as was the case for
HiPCo SWNT in UPR.14 Somewhat surprisingly, neither the
dierences in chirality (and therefore diameter distribution)
8645

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

Figure 4. ColeCole plot of G versus G variation with angular frequency at dierent SWNT concentrations. Concentrations: 0.05 (0.005), 0.095
(0.005), 0.135 ( 0.005), 0.18 ( 0.01) and 0.23 ( 0.02) vol %. Error: < 12 %.

Figure 5. Master curves for SWNT-UPR dispersions superimposed using Gc and c as shift factors. (a) CG200, (b) CG300, (c) SG65, (d) SG65i.
Concentrations: 0.05 (0.005), 0.095 (0.005), 0.135 ( 0.005), 0.18 ( 0.01), 0.23 ( 0.02) vol %. Error: < 12 %.

Nonetheless, the similar values of obtained for the dierent


SWNT dispersions indicate a similar type of attraction and
network structure. Making the proportionality an equality (Go
= A( c)) and calculating the proportionality constant
enables comparison of the relative nanotubenanotube
attraction in each system. As shown by Prasad et al. in their
study of model colloid systems, higher values of A indicate
greater interparticle attraction.29 From Figure 7e,f, the A values
for CG200, CG300, SG65, and SG65i were 3300, 6470, 1780,

and 5920, respectively; these dierences in attraction result in


the dierences in viscoelastic properties. It is hypothesized that
the presence of more non-SWNT carbon lowered the total
attractive potential in the SG65 (and to a lesser extent in
CG200); non-SWNT carbon have lower interparticle attractions with each other and mediate SWNT-SWNT attraction by
disrupting SWNT packing.
Following the work of Khalkhal et al.7 and Urena-Benavidas
et al.10 for MWNT-epoxy and MWNT-UPR, the fractal
8646

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

Figure 6. Damping factor tan as a function of angular frequency (a) CG200, (b) CG300, (c) SG65, (d) SG65i; 0.050 (0.005), 0.095 (0.005),
0.135 ( 0.005), 0.18 ( 0.01) and 0.23 ( 0.02) vol %. Error bars < 12 %.

observed for the SWNT types indicate faster aggregating


systems; this was also observed for HiPCo SWNT-UPR and
MWNT-UPR.10,14 Intuitively, both the presence of non-SWNT
carbon and diameter distribution would be expected to aect df.
Impurities could both slow aggregation and disrupt uniform
packing; similarly, a more uniform diameter distribution might
facilitate faster, tighter packing. However, the similarity of df for
CG300 and SG65 suggest other factors such as the total
amount of solids and size of impurities may be involved in
microstructure formation. Similarly, CG300 and SG65 had the
highest values of m, indicating that they had the softest ocs;
the m values for CG200 and SG65i were in the range
considered characteristic of rigid ocs. For all SWNT types, the
values were consistent with both inter- and intraoc links
contributing to elasticity. Similar values were observed for
HiPCO SWNT-UPR.10,14 For CG200, intraoc links were
more dominant than for the other SWNT grades.
A nal way to probe microstructural dierences is through
the relaxation dynamics. The shape of the G versus curve
and the crossover point c where G = G = Gc are a reection
of the systems relaxation modes.4,33,34 Frequencies above the
crossover frequency c probe the short-range dynamics of small
length scales comparable to entanglement length. The
microstructure of solid particles restrains polymer chain longrange mobility, and the overall system shows slow relaxation.
Therefore, below c longer range dynamics are probed, and G
> G due to the elastic behavior of the nanomaterial
microstructure. With increasing concentration, the solid
microstructure becomes more compact which leads to a
attening of the G curve resulting from longer time scale
relaxation. All SWNT-UPR dispersions showed a decrease in G
slope and increase in c with increasing concentration. CG200,
CG300, and SG65i consistently displayed higher c (tan = 1)
compared to SG65-UPR at similar SWNT concentration
(Figure 6 and Table 5). This means the microstructure of the
CG200, CG300, and SG65i dispersions induces nanotube

dimensions comprising the network were calculated from


scaling relations for low frequency plateau modulus and critical
strain using methods developed by Shih et al. and Wu and
Morbidelli.30,31 The equations for the scaling relations are the
following:
G B /(d d f )

(1)

c (d B 1)/(d d f )

(2)

where df is the fractal dimension and d is the system


dimensionality.7,10 The value of df can be calculated by
equating the exponents of the two scaling relations. According
to Wu and Morbidelli, B = (d 2) + (2 + x)(1 ), where 1
x 1.3 is the backbone fractal dimension, d = 3 for a threedimensional network, and ranges from 0 to 1 and denes the
interoc and intraoc strength.7,31 For = 0, the interoc links
are stronger than the intraoc links, while for = 1 the intraoc
strength dominates.31 For rigid aggregates, the interaction
potential of the particles is noncentral and m = 1/B and is
estimated as being between 0.23 and 0.29.7 For soft aggregates
(or ocs) the interaction is central and depends on the distance
between particle centers, and the internal structure does not
respond elastically to small deformations. In this case, m is
approximately 0.40.5.32 It should be noted that while this
method can provide a wealth of information about dispersion
microstructure, the results are strongly dependent on the
choice of the critical strain c. Underestimating c can lead to
nonphysical negative values of the fractal dimension df. Direct
quantitative comparisons should only be done for systems
where the same criteria was used to determine c. In this work,
c represents the strain at which G decreased by 2% during the
amplitude sweep; the corresponding results are shown in Table
4.
Fractal dimensions of 1.7 < df < 1.8 have been associated
with fast aggregating systems while 2.0 < df < 2.2 have been
associated with slower aggregating ocs.10 The df values
8647

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

Figure 7. Divergence modulus vs concentration for (a) CG200, (b) CG300, (c) SG65 and (d) SG65i. Scaling relation Go ( c) for (e)
CG200, CG300 and (f) SG65, SG65i (from the gure y = Go and x = c).

dominated relaxation dynamics at a lower concentration than


SG65. This further supports dierences in the microstructural
physical interactions due to non-SWNT content. While there
were dierences between SG65 and other SWNTs in
approaching solid dominated relaxation mode, none of the
SWNT systems showed a second relaxation mode of the solid
network at the lower limit of frequency as observed in the case
of SWNTPDMS dispersion characterized by Marceau et al.33
Hassanabadi et al. showed that CNT showed higher elasticity
and slower relaxation dynamics than nanoclay in ethylene vinyl
acetate; this was the result of a stronger network formed by
higher internanotube attraction and a higher degree of
connement of polymer by the CNT network.35 Similar
premises can be used to explain the greater solid-like behavior
of the higher purity SWNT in UPR compared to the SG65UPR dispersions. We propose that the network strength can
also be quantied based on the gel stiness S using the G() =
S(1 n) cos(n/2)n relation developed by Mours and
Winter.36 While S provides similar information as G(),
characterizing dispersions in terms of a single stiness
parameter is more intuitive. The stiness parameter has been
predominantly used for comparing chemically cross-linked gels
and physical gels of rod-like virus particle suspensions.3638 We
are proposing this parameter could also be used to characterize
stiness of systems in the vicinity of percolation. Based on the

Table 4. Fractal Structure Parameters for SWNT-UPR


Dispersions
SWNT
type

B/
(d df)

(d B
1)/(d df)

df

CG200
CG300
SG65
SG65i

2.26
2.48
2.18
2.28

1.060
0.948
0.702
0.956

1.33
1.69
1.65
1.48

3.77
3.25
2.94
3.27

0.26
0.31
0.34
0.29

0.080.16
0.250.32
0.350.42
0.240.31

Table 5. Crossover Frequencies and Stiness Factors


CG200

CG300

vol %

c (rad/s)

vol %

c (rad/s)

0.09
0.13
0.17
0.22

10.0
25.1
39.8
99.8
SG65

94
250
290
670

0.10
0.14
0.17
0.20

15.8
25.1
39.8
63.0
SG65i

150
280
550
820

vol %

c (rad/s)

vol %

c (rad/s)

0.10
0.14
0.19
0.25

2.51
10.0
15.8
25.1

65
110
210
340

0.10
0.14
0.17
0.21

15.8
25.1
39.8
63.0

140
295
474
760

8648

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules

(3) Chatterjee, T.; Krishnamoorti, R. Steady Shear Response of


Carbon Nanotube Networks Dispersed in Poly (Ethylene Oxide).
Macromolecules 2008, 41 (14), 53335338.
(4) Du, F.; Scogna, R. C.; Zhou, W.; Brand, S.; Fischer, J. E.; Winey,
K. I. Nanotube Networks in Polymer Nanocomposites: Rheology and
Electrical Conductivity. Macromolecules 2004, 37 (24), 90489055.
(5) Fan, Z.; Advani, S. G. Rheology of Multiwall Carbon Nanotube
Suspensions. J. Rheol. 2007, 51 (4), 585604.
(6) Hobbie, E.; Fry, D. Rheology of Concentrated Carbon Nanotube
Suspensions. J. Chem. Phys. 2007, 126 (12), 124907.
(7) Khalkhal, F.; Carreau, P. J. Scaling Behavior of the Elastic
Properties of Non-Dilute MWCNTEpoxy Suspensions. Rheol. Acta
2011, 50 (910), 717728.
(8) Mitchell, C. A.; Bahr, J. L.; Arepalli, S.; Tour, J. M.;
Krishnamoorti, R. Dispersion of Functionalized Carbon Nanotubes
in Polystyrene. Macromolecules 2002, 35 (23), 88258830.
(9) Song, Y. S.; Youn, J. R. Influence of Dispersion States of Carbon
Nanotubes on Physical Properties of Epoxy Nanocomposites. Carbon
2005, 43 (7), 13781385.
(10) Urena-Benavides, E. E.; Kayatin, M. J.; Davis, V. A. Dispersion
and Rheology of Multiwalled Carbon Nanotubes in Unsaturated
Polyester Resin. Macromolecules 2013, 46 (4), 16421650.
(11) Ao, G.; Khripin, C. Y.; Zheng, M. DNA-Controlled Partition of
Carbon Nanotubes in Polymer Aqueous Two-Phase Systems. J. Am.
Chem. Soc. 2014, 136 (29), 1038310392.
(12) Yakobson, B. I.; Couchman, L. S. Carbon Nanotubes:Superamolecular Mechanics. In Dekker Encyclopedia of Nanoscience and
Nanotechnology; Schwarz, J. A., Lyshevski, E. E., Putyera, K., Contescu,
C. I., Eds.; Marcel Dekker: New York, 2004; Vol. 1, pp 587601.
(13) Menzer, K.; Krause, B.; Boldt, R.; Kretzschmar, B.; Weidisch, R.;
Potschke, P. Percolation Behaviour of Multiwalled Carbon Nanotubes
of Altered Length and Primary Agglomerate Morphology in Melt
Mixed Isotactic Polypropylene-Based Composites. Compos. Sci.
Technol. 2011, 71 (16), 19361943.
(14) Kayatin, M. J.; Davis, V. A. Viscoelasticity and Shear Stability of
Single-Walled Carbon Nanotube/Unsaturated Polyester Resin Dispersions. Macromolecules 2009, 42 (17), 66246632.
(15) Jansen, R.; Wallis, P. Manufacturing, Characterization and Use
of Single Walled Carbon Nanotubes. Mater. Matters 2009, 4, 2328.
(16) Nepal, D.; Balasubramanian, S.; Simonian, A. L.; Davis, V. A.
Strong Antimicrobial Coatings: Single-Walled Carbon Nanotubes
Armored with Biopolymers. Nano Lett. 2008, 8 (7), 18961901.
(17) Ziegler, K. J.; Rauwald, U.; Gu, Z.; Liang, F.; Billups, W.; Hauge,
R. H.; Smalley, R. E. Statistically Accurate Length Measurements of
Single-Walled Carbon Nanotubes. J. Nanosci. Nanotechnol. 2007, 7 (8),
29172921.
(18) Resasco, D.; Alvarez, W.; Pompeo, F.; Balzano, L.; Herrera, J.;
Kitiyanan, B.; Borgna, A. A Scalable Process for Production of SingleWalled Carbon Nanotubes (SWNTS) by Catalytic Disproportionation
of Co on a Solid Catalyst. J. Nanopart. Res. 2002, 4 (12), 131136.
(19) Chatterjee, T.; Krishnamoorti, R. Dynamic Consequences of the
Fractal Network of Nanotube-Poly(Ethylene Oxide) Nanocomposites.
Phys. Rev. E 2007, 75 (5), 050403.
(20) Kharchenko, S. B.; Douglas, J. F.; Obrzut, J.; Grulke, E. A.;
Migler, K. B. Flow-Induced Properties of Nanotube-Filled Polymer
Materials. Nat. Mater. 2004, 3 (8), 564568.
(21) Filippone, G.; Salzano de Luna, M. A Unifying Approach for the
Linear Viscoelasticity of Polymer Nanocomposites. Macromolecules
2012, 45 (21), 88538860.
(22) Macosko, C. W.; Larson, R. G. Rheology: Principles, Measurements, and Applications; Wiley-VCH: Weinheim, 1994.
(23) Chen, D.; Chen, K.; Hough, L.; Islam, M.; Yodh, A. Rheology of
Carbon Nanotube Networks During Gelation. Macromolecules 2010,
43 (4), 20482053.
(24) Trappe, V.; Weitz, D. Scaling of the Viscoelasticity of Weakly
Attractive Particles. Phys. Rev. Lett. 2000, 85 (2), 449.
(25) Pogodina, N. V.; Winter, H. H. Polypropylene Crystallization as
a Physical Gelation Process. Macromolecules 1998, 31 (23), 8164
8172.

stiness factor values shown in Table 5, the network stiness


values are similar for CG300 and SG65i and decrease with
increasing non-SWNT carbon content. This highlights how the
higher attractive forces in the high purity dispersions aect
viscoelasticity.

CONCLUDING REMARKS
It has long been recognized that dierences between SWNT
products and even batch to batch variation may cause
dierences in dispersion state and nal composite properties.
However, it has only recently been possible to systematically
compare the eects of chirality distribution and purity. This
work shows that in the case of dispersion in UPR the presence
of non-SWNT carbon has a much greater eect on
viscoelasticity and microstructure than the dierences in
chirality distribution between current commercial grades. It
also provides a framework for comprehensive comparison of
dispersions in terms of measured rheological properties,
percolation threshold, nanomaterialpolymer interaction,
occulation kinetics, fractal dimensions, oc linkage strength,
relaxation dynamics, and stiness. The methodologies used in
this paper should be broadly applicable to understanding if
chirality distribution aects the microstructure of dispersions in
melts, other thermosets, and polymer solutions. The methodologies should also be broadly applicable to other dispersions
exhibiting colloidal rod behavior including, but not limited to
future commercial nanotube grades. While, on the one hand,
there are ongoing eorts to produce even higher quality SWNT
with narrower chirality distributions and higher purities, a
growing number of researchers are exploring production of
lower cost nanotubes from waste carbon sources. Therefore,
tools for systematically comparing dispersion properties will be
of increasing importance.

ASSOCIATED CONTENT

* Supporting Information
S

The Supporting Information is available free of charge on the


ACS Publications website at DOI: 10.1021/acs.macromol.5b00870.
Additional nanotube characterization data, rheology
plots, model parameters, and shift factors (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail davisva@auburn.edu (V.A.D.).


Notes

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The authors acknowledge NSF EPSCOR RII Award 1158862
and the Department of Education Graduate Awards in Areas of
National Need (GAANN) P200A120228.

REFERENCES

(1) Girifalco, L.; Hodak, M.; Lee, R. S. Carbon Nanotubes,


Buckyballs, Ropes, and a Universal Graphitic Potential. Phys. Rev. B:
Condens. Matter Mater. Phys. 2000, 62 (19), 13104.
(2) OConnell, M. J.; Bachilo, S. M.; Huffman, C. B.; Moore, V. C.;
Strano, M. S.; Haroz, E. H.; Rialon, K. L.; Boul, P. J.; Noon, W. H.;
Kittrell, C. Band Gap Fluorescence from Individual Single-Walled
Carbon Nanotubes. Science 2002, 297 (5581), 593596.
8649

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Article

Macromolecules
(26) Romasanta, L.; Lopez-Manchado, M.; Verdejo, R. The Role of
Carbon Nanotubes in Both Physical and Chemical LiquidSolid
Transition of Polydimethylsiloxane. Eur. Polym. J. 2013, 49 (6), 1373
1380.
(27) Arbabi, S.; Sahimi, M. Mechanics of Disordered Solids. I.
Percolation on Elastic Networks with Central Forces. Phys. Rev. B:
Condens. Matter Mater. Phys. 1993, 47 (2), 695.
(28) Surve, M.; Pryamitsyn, V.; Ganesan, V. Universality in Structure
and Elasticity of Polymer-Nanoparticle Gels. Phys. Rev. Lett. 2006, 96
(17), 177805.
(29) Prasad, V.; Trappe, V.; Dinsmore, A.; Segre, P.; Cipelletti, L.;
Weitz, D. Rideal Lecture Universal Features of the Fluid to Solid
Transition for Attractive Colloidal Particles. Faraday Discuss. 2003,
123, 112.
(30) Shih, W.-H.; Shih, W. Y.; Kim, S.-I.; Liu, J.; Aksay, I. A. Scaling
Behavior of the Elastic Properties of Colloidal Gels. Phys. Rev. A: At.,
Mol., Opt. Phys. 1990, 42 (8), 4772.
(31) Wu, H.; Morbidelli, M. A Model Relating Structure of Colloidal
Gels to Their Elastic Properties. Langmuir 2001, 17 (4), 10301036.
(32) Potanin, A. A. On the Computer Simulation of the Deformation
and Breakup of Colloidal Aggregates in Shear Flow. J. Colloid Interface
Sci. 1993, 157 (2), 399410.
(33) Marceau, S.; Dubois, P.; Fulchiron, R.; Cassagnau, P.
Viscoelasticity of Brownian Carbon Nanotubes in PDMS Semidilute
Regime. Macromolecules 2009, 42 (5), 14331438.
(34) Pryamitsyn, V.; Ganesan, V. Origins of Linear Viscoelastic
Behavior of Polymer-Nanoparticle Composites. Macromolecules 2006,
39 (2), 844856.
(35) Hassanabadi, H. M.; Wilhelm, M.; Rodrigue, D. A Rheological
Criterion to Determine the Percolation Threshold in Polymer NanoComposites. Rheol. Acta 2014, 53, 869.
(36) Mours, M.; Winter, H. H. Relaxation Patterns of Nearly Critical
Gels. Macromolecules 1996, 29 (22), 72217229.
(37) Reddy, N. K.; Zhang, Z.; Lettinga, M. P.; Dhont, J. K.; Vermant,
J. Probing Structure in Colloidal Gels of Thermoreversible Rodlike
Virus Particles: Rheology and Scattering. J. Rheol. 2012, 56 (5), 1153
1174.
(38) Sarvi, A.; Sundararaj, U. Electrical Permittivity and Electrical
Conductivity of Multiwall Carbon Nanotube-Polyaniline (MWCNTPANI) Core-Shell Nanofibers and MWCNT-PANI/Polystyrene
Composites. Macromol. Mater. Eng. 2014, 299 (8), 10131020.

8650

DOI: 10.1021/acs.macromol.5b00870
Macromolecules 2015, 48, 86418650

Вам также может понравиться