Sei sulla pagina 1di 4

Materials Letters 62 (2008) 34933496

Contents lists available at ScienceDirect

Materials Letters
j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / m a t l e t

Quantitative study of the dispersion degree in carbon nanober/polymer and carbon


nanotube/polymer nanocomposites
Z.P. Luo a,, J.H. Koo b
a
b

Microscopy and Imaging Center, Biological Sciences Building West, Texas A&M University, College Station, TX 77843-2257, USA
Department of Mechanical Engineering, The University of Texas at Austin, Austin, TX 78712-0292, USA

a r t i c l e

i n f o

Article history:
Received 4 February 2008
Accepted 4 March 2008
Available online 18 March 2008
Keywords:
Electron microscopy
Microstructure
Nanocomposites
Polymers

a b s t r a c t
Quantitative measurements of the ller dispersion degree of carbon nanober (CNF) and nanotube (CNT)
reinforced polymer nanocomposites have been made by transmission electron microscopy. Samples were
prepared by either high-shear mixing or twin-screw extrusion processing. It was found that the ller
dispersion degree was largely inuenced by the ller size. As the ller dimension became smaller, the
dispersion parameter D0.1 largely decreased as quantied, which demonstrated the challenges associated with
improving the dispersion of smaller llers. This work provided a method to quantitatively compare the
dispersion degrees of CNF/CNT polymer nanocomposites.
2008 Elsevier B.V. All rights reserved.

1. Introduction
Carbon nanobers (CNFs), with diameter of several hundreds of
nanometers, and carbon nanotubes (CNTs), with even smaller
dimensions from several to tens of nanometers, are increasingly
favored as llers in advanced polymer nanocomposites because these
llers signicantly improve the performance of the composites. It is
generally accepted that the composite properties are dominated by
the dispersion degree of the llers. Tibbetts and McHugh found that
reducing the ber clump size could signicantly improve the CNF/
polymer composites properties [1]. An early work on the CNF/polymer
showed a peak tensile strength of 69 MPa at 5 wt.% CNF, but then the
strength decreased at higher contents [2], which might be related to
the poor dispersion. The strain levels were found to be 67% and 44% at
5 and 10 wt.% contents, respectively. Later, a well-dispersed CNF/
polymer composite exhibited a strain as high as 1300% at 8 wt.% CNF
content [3]. Patton et al. [4] found that, with increasing of the ber
content, the strength and modulus increased to peak levels at 15.5 vol.
% and 19.2 vol.%, respectively, and afterwards they decreased. The
work by Kuriger et al. showed that the composite strength increased
with ber volume fraction and the degree of ber alignment, while at
higher ber volume fractions, the composite strength was much lower
than predicted by theory, which might be related to poor dispersion
caused by the incomplete wetting and inltration of the CNFs [5].
Because the composite properties are strongly correlated to their
microstructure, detailed microstructural studies are essential to
understanding the microstructureproperty relationship [610].

Corresponding author. Tel.: +1 979 845 1129; fax: +1 979 847 8933.
E-mail address: luo@mic.tamu.edu (Z.P. Luo).
0167-577X/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.matlet.2008.03.010

Andrews et al. studied the dispersion of CNF in polished sections of


composites using optical microscopy [11]. They assigned the dispersion degree from 1 to 10, based upon the uniformity of ber
distribution across the specimen, and in particular upon the frequency
of occurrence of agglomerates.
The transmission electron microscopy (TEM) has provided a
resolution normally around the several angstroms level that enables
the studies of these composites with nanollers. However, currently
the TEM is mainly used as a tool to compare the dispersion degree
visually. In this work, we present an unambiguous dispersion quantication of the CNF/CNT composites using TEM.
2. Experimental procedures
Four CNF/CNT llers were used: (1) PR-19-PS (100200 nm
diameter) and PR-24-PS (60150 nm diameter), both with 30100 m
length, were manufactured by Applied Sciences, Inc.; (2) CM-95 multiwall nanotubes (MWNTs), with 1015 nm diameter and 1020 m
length, were manufactured by ILJIN Nanotech Co., Ltd.; and (3) smalldiameter nanotubes (SDNTs), with less than 3.5 nm diameter and about
several m length, were manufactured by Carbon Nanotechnologies, Inc.
The polymer matrices used in this study were: (1) thermoplastics of
polyamide PA11 (or nylon 11), and polyamide PA12 (or nylon 12);
(2) thermoplastic elastomer (TPE) of polyester polycaprolactone
(pellethane 2102 series) 2102A, TPE X1180, and thermoplastic polyurethane (TPU) elastomer; and (3) thermosetts of cyanate ester PT30
and PT15, epoxy, and phenolic resin SC1008. The thermosetting-based
nanocomposites were prepared by high-shear mixing, and the thermoplastic-based and thermoplastic elastomer-based nanocomposites were
compounded using a commercial twin-screw extruder. The TEM
specimens were prepared by either ion mill for hard material, or

3494

Z.P. Luo, J.H. Koo / Materials Letters 62 (2008) 34933496

ultramicrotomy for soft material. The TEM analysis was performed by a


JEOL 2010 at 200 kV voltage.
The dispersion measurements of the CNFs/CNTs polymer nanocomposites were made by applying a methodology proposed by us
recently [12]. First, a randomly taken representative TEM image was
normally divided into 1020 grid lines along the horizontal and
vertical directions, and then the free-path spacing between the llers
were measured using a freeware ImageJ [13] after magnication
calibration using standards of 6H SiC lattice fringes [14] and cross-line
grating replica. The dispersion parameter, D0.1, is deed as the
probability of the free-path distance distribution in the range of
0.91.1, where is the mean spacing. Usually the free-path distance
distribution obeys a lognormal distribution model, in which a case D0.1
is formularized as [12]:
D0:1 1:1539  102 7:5933  102 A=r 6:6838
 104 A=r2 1:9169  104 A=r3 3:9201
 106 A=r4 :

Here, is the standard deviation. Therefore, once the spacing data


are measured and the data sample mean
q
P
s
2 =N  1
i xi  x

xi =N

and data sample

are obtained, the dispersion


standard deviation
quantity D0.1 can be obtained by using the ratio x/s to substitute / in
the Eq. (1). Here, x and s are used to estimate and , respectively.
3. Results and discussion
In this work, the dispersion degrees of four llers in different matrices are examined, as
listed in Table 1. The PR-19-PS and PR-24-PS CNFs exhibit good dispersions. Between them,
the PR-19-PS (with 150 nm average diameter) has slightly higher dispersions than PR-24-PS
(with 105 nm average diameter). However, as compared to CNFs, the MWNTs and SDNTs
have much lower dispersions. Between them, the SDNT, with the smallest diameter (less
than 3.5 nm), has the lowest dispersion. Typical examples are given as follows.
A representative microstructure of PA11 with 5 wt.% PR-19-PS CNF is shown in Fig.1(a).
The CNFs are seen well dispersed into the matrix without any large clumps. This entire area
is divided by 10 10 grid lines along the horizontal and vertical directions, respectively, and
then the free-path spacing between the bers along the grid lines are measured, with
sampling number of N = 301. The spacing histogram is plotted in Fig. 1(b). A lognormal
distribution [12] is imposed, which ts well the measured frequency curve, and thus the
dispersion degree is calculated by the Eq. (1). Instead, a normal distribution only gives
symmetrical curve along the peak position. The measured average spacing is x = 326.5 nm,
with standard deviation of s = 261.3 nm, and thus x/s = 1.2495. Hence, D0.1 = 10.7% by using
x/s to replace /. Therefore, about 10.7% spacing data are in the range of 293.9359.2 nm
that corresponds to the range of 0.9x1.1x.
However in the sample of MWNT with smaller size ller, the dispersion level is
apparently decreased. A typical microstructure of PT15 matrix with 1 wt.% MWNTs is
shown is Fig. 2(a), with an enlargement from the framed area shown in Fig. 2(b) for
details. It is seen that the llers form certain aggregations, leaving some areas unlled.
The measured histogram from Fig. 2(a) along 10 10 grid lines is shown in Fig. 2(c),
based on the measurements with sampling number of N = 543. Because of the existence
of some unlled areas, the range of the spacing data becomes greater, from 01800 nm

Table 1
Dispersion quantication results
Filler

Composites

D0.1

PR-19-PS (100200 nm)

Epoxy + 5 wt.%PR-19-PS
PA11 + 5 wt.%PR-19-PS
PA12 + 5 wt.%PR-19-PS
TPE X1180 + 2.5 wt.%PR-19-PS
TPU + 5 wt.%PR-19-PS
2102A + 15 wt.%PR-24-PS
Epoxy + 5 wt.%PR-24-PS
PT30 + 0.5 wt.%PR-24-PS
SC1008 + 10 wt.%PR-24-PS
TPU + 5 wt.%PR-24-PS
PT15 + 1 wt.%MWNT
PT15 + 5 wt.%MWNT
PT15 + 1 wt.%SDNT
PT15 + 5 wt.%SDNT

9.6
10.7
10.8
9.2
10.5
9.8
8.8
9.0
10.0
9.9
5.5
6.3
3.3
4.3

PR-24-PS (60150 nm)

MWNT (1015 nm)


SDNT (b 3.5 nm)

Fig. 1. (a) Microstructure of PA11 with 5 wt.% PR-19-PS CNF composite; (b) histogram of the
ller free-path spacing. Arrowhead indicates the mean spacing position. Measurement
statistics and quantication results are labeled.

as shown in Fig. 2(c). A magnied part in the range of 0400 nm is shown in Fig. 2(d).
The appearance of a high frequency peak in a short distance in Fig. 2(d) is an indication
of the ller clustering. The average spacing is measured as x = 102.0 nm with standard
deviation of s = 184.8 nm, x/s = 0.5519, and thus D0.1 = 5.5% according to Eq. (1), which is
signicantly lower than the dispersion degree of D0.1 = 10.7% for PR-19-PS in Fig. 1.
When the ller size becomes smaller than MWNTs, as shown in Fig. 3(a), the PT15
with 1 wt.% SDNTs presents even severe tangling. An enlarged image is shown in Fig. 3
(b). The histogram from the spacing measurements from 10 10 grid lines is shown in
the histogram in Fig. 3(c), with x = 140.1 nm, s = 565.8 nm, x/s = 0.2476, and thus
D0.1 = 3.3%, which is lower than the dispersion of 5.5% for 1 wt.% MWNT in Fig. 2. An
enlargement in the 0200 nm range of Fig. 3(c) is shown in Fig. 3(d).
The tangling of the llers in the polymer may be caused by the intrinsic van der
Waals forces during the composite fabrication process. As compared to the CNFs, such
intrinsic forces of CNTs seem to be stronger than the CNFs. In fact, Gojny et al. [15]
noticed that MWNTs exhibited better dispersion than the double-wall nanotubes in the
epoxy matrix. In recent years, there have developed several ways to achieve better
dispersions of CNTs integrated into the polymer matrices, including optimum physical
blending (high-shear mixing and ultrasonication) [16,17], surfactant-assisted processing [18], in-situ polymerization [19], and functionalization [20]. However, how to
achieve better dispersion of the smaller nanotubes in our systems is not within the
scope of this paper.
Using the TEM methodology presented in this work, the dispersion degree D0.1 can
be quantitatively compared for different samples. As the parameter is only related to
ller free-path spacing distribution, it is independent of the CNF/CNT ller shape, size,
or content.

Acknowledgements
This work was supported by the Air Force Ofce of Scientic Research
(AFOSR), Arlington, VA. The authors thank Rick Littleton for technical
assistance in the TEM sample preparation, Dr. Mike Pendleton for

Z.P. Luo, J.H. Koo / Materials Letters 62 (2008) 34933496

3495

Fig. 2. (a) Microstructure of PT15 with 1 wt.% MWNT composite; (b) enlargement from the framed area in (a); (c) histogram of the ller free-path spacing measured from (a);
(d) enlarged histogram from (c) showing the short spacing range.

Fig. 3. (a) Microstructure of PT15 with 1 wt.% SDNT composite; (b) enlargement from the framed area in (a); (c) histogram of the ller free-path spacing measured from (a);
(d) enlarged histogram from (c) showing the short spacing range.

3496

Z.P. Luo, J.H. Koo / Materials Letters 62 (2008) 34933496

assistance in the manuscript preparation, and Dr. Andreas Holzenburg


for his encouragement to this research.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]

G.G. Tibbetts, J.J. McHugh, J. Mater. Res. 14 (1999) 28712880.


K. Lozano, E.V. Barrera, J. Appl. Polym. Sci. 79 (2000) 125133.
K. Lozano, S.Y. Yang, R.E. Jones, Carbon 42 (2004) 23292331.
R.D. Patton, C.U. Pittman, L. Wang, J.R. Hill, Composites A30 (1999) 10811091.
R.J. Kuriger, M.K. Alam, D.P. Anderson, J. Mater. Res. 16 (2001) 226232.
P. Ptschke, A.R. Bhattacharyya, A. Janke, Eur. Polym. J. 40 (2004) 137148.
F.H. Gojny, M.H.G. Wichmann, U. Kpke, B. Fiedler, K. Schulte, Compos. Sci. Technol.
64 (2004) 23632371.
[8] H.G. Chae, T.V. Sreekumar, T. Uchida, S. Kumar, Polymer 46 (2005) 1092510935.
[9] M.M. Hasan, Y.X. Zhou, S. Jeelani, Mater. Lett. 61 (2007) 11341136.

[10] K.P. Ryan, M. Cadek, V. Nicolosi, D. Blond, M. Ruether, G. Armstrong, H. Swan, A.


Fonseca, J.B. Nagy, W.K. Maser, W.J. Blau, J.N. Coleman, Compos. Sci. Technol. 67
(2007) 16401649.
[11] R. Andrews, D. Jacques, M. Minot, T. Rantell, Macromol. Mater. Eng. 287 (2002)
395403.
[12] Z.P. Luo, J.H. Koo, J. Microsc. 225 (2007) 118125.
[13] M.D. Abramoff, P.J. Magelhaes, S.J. Ram, Biophoton. Int. 11 (2004) 3642.
[14] Z.P. Luo, Acta Mater. 54 (2006) 4758.
[15] F.H. Gojny, M.H.G. Wichmann, B. Fiedler, K. Schulte, Compos. Sci. Technol. 65 (2005)
23002313.
[16] E.T. Thostenson, T.W. Chou, J. Phys. D: Appl. Phys. 35 (2002) L77L80.
[17] P. Ptschke, A.R. Bhattacharyya, A. Janke, Carbon 42 (2004) 965969.
[18] X.Y. Gong, J. Liu, S. Baskaran, R.D. Voise, J.S. Young, Chem. Mater. 12 (2000)
10491052.
[19] B.Z. Tang, H.Y. Xu, Macromolecules 32 (1999) 25692576.
[20] A. Oki, L. Adams, V. Khabashesku, Y. Edigin, P. Biney, Z.P. Luo, Mater. Lett. 62 (2008)
918922.

Potrebbero piacerti anche