You are on page 1of 6

Nanotechnology

You may also like


- Simulations of guiding of low-energy ions
Creep mitigation in composites using carbon through a single nanocapillary in insulating
materials
nanotube additives Shi-Dong Liu, , Yong-Tao Zhao et al.

- Effect of size quantization and quantum


capacitance on the threshold voltage of a
To cite this article: W Zhang et al 2007 Nanotechnology 18 185703 2D nanoscale dual gate MOSFET
Abhinav Sundar and Niladri Sarkar

- Recent progress in near-field


nanolithography using light interactions
with colloidal particles: from nanospheres
View the article online for updates and enhancements. to three-dimensional nanostructures
Xu A Zhang, I-Te Chen and Chih-Hao
Chang

This content was downloaded from IP address 193.137.97.161 on 14/12/2022 at 16:25


IOP PUBLISHING NANOTECHNOLOGY
Nanotechnology 18 (2007) 185703 (5pp) doi:10.1088/0957-4484/18/18/185703

Creep mitigation in composites using


carbon nanotube additives
W Zhang1 , A Joshi2 , Z Wang1 , R S Kane2 and N Koratkar1,3
1
Department of Mechanical Aerospace and Nuclear Engineering, Rensselaer Polytechnic
Institute, 110 8th Street, Troy, NY 12180, USA
2
Department of Chemical and Biological Engineering, Rensselaer Polytechnic Institute,
110 8th Street, Troy, NY 12180, USA

E-mail: koratn@rpi.edu

Received 19 February 2007, in final form 14 March 2007


Published 5 April 2007
Online at stacks.iop.org/Nano/18/185703
Abstract
A major limitation of thermosetting epoxy based polymeric materials in
long-term structural applications is mechanical creep. Here it is demonstrated
that single-walled carbon nanotube additives in low weight fractions
(0.1–0.25%) are effective in limiting the load-induced re-orientation of epoxy
chains, resulting in a significant slowing of the creep response. Nanotube
additives could therefore be the key enabler for the long-term
higher-temperature application of polymeric structures which would
otherwise fail by excessive creep deformation.
(Some figures in this article are in colour only in the electronic version)

Over the past decade, carbon nanotube composites [1–7] have bonds. The chemical bonding is strong and directional along
been researched extensively for their thermal, mechanical, the chains, but adjacent chains are bonded by weak Van der
electrical and optical properties. The mechanical proper- Waals forces. Under persistent external loading, the molecular
ties that have been investigated include strength [8, 9], stiff- chains are stretched and re-oriented to carry the load, resulting
ness [10, 11], buckling resistance [12], wear [13] and tough- in mechanical creep. The creep deformation is strongly
ness [14] as well as viscoelastic properties [15–19] such as dependent on the stress level, the temperature and the micro-
loss factor and loss modulus. However, as far as we are aware, structure of the materials. Some recent studies [20, 21] have
there are no reports in the literature on creep characterization shown that the addition of nanoparticles into a polymer matrix
of carbon nanotube composite materials. Creep is the time- enhances its creep resistance at various stress and temperature
dependent continued deformation of a structure under a persis- levels. In another study [22], the creep behaviour of
tent load that is usually significantly lower than the material’s nanocomposites that incorporate layered silicate in the matrix
yield stress. Materials with high creep resistance are critical is investigated. A comprehensive review of carbon nanotube
in long-term structural applications to ensure dimensional sta- epoxy composites, including stiffness and fracture toughness
bility and load-carrying capability. In this paper a series of characterization, is described in [27]. However, no systematic
creep characterization tests are performed for a carbon nan- study of the tensile creep behaviour of carbon nanotube filled
otube epoxy composite system, and a significant reduction in polymer composites is available in the literature. In the
creep is demonstrated with a relatively small weight fraction
following sections we describe the protocols used to fabricate
(∼0.1%) of nanotube additives. To our knowledge, this is the
the nano-composite test coupons and the experiments that were
first reported demonstration of creep reduction in polymer sys-
performed to characterize the creep behaviour of the nanotube–
tems by the use of carbon nanotube additives.
epoxy composites.
Thermosetting epoxy polymers were chosen for this study
The procedure used to disperse single-walled carbon nan-
since they are widely used in a variety of applications such as
paints, coatings, adhesives, industrial tooling and composites, otubes (SWNT) in the epoxy system is shown schematically
as well as in the semiconductor and electronics packaging in figure 1(A) and described in the experimental section (ap-
industries. Epoxies consist of long molecules that are cross- pendix). The samples were tested using an MTS-858 servo-
linked and contain chains of atoms held together by covalent hydraulic test system operating in constant-force mode. The
force is measured by the load cell of the MTS system and an in-
3 Author to whom any correspondence should be addressed.
ternal feedback circuit is used to maintain constant stress in the

0957-4484/07/185703+05$30.00 1 © 2007 IOP Publishing Ltd Printed in the UK


Nanotechnology 18 (2007) 185703 W Zhang et al

A C

D
B

Figure 1. Creep characterization of nano-composite and pure epoxy samples. (A) Schematic showing the fabrication protocol used to disperse
the single-walled carbon nanotube additives into the epoxy matrix. The exact same fabrication steps were also repeated to prepare the baseline
epoxy samples without the nanotube additives. (B) Measured creep strain as a function of time. The room-temperature creep response of the
0.1 wt% nano-composite is significantly lower than the pure epoxy. After 50 h of continuous testing at 10 MPa stress, the nano-composite
shows a reduction of over 30% in the creep deformation. (C) Results for percentage reduction in creep strain for nano-composite compared to
pure epoxy sample after 50 h of testing at 10 MPa stress for different weight fractions of nanotube additives. The weight fractions for
minimum creep response lie in the 0.1–0.25% range. (D) Scanning electron microscopy characterization of microstructure of the 0.1 wt% (top
images) and 1 wt% (bottom images) samples. The 1 wt% nano-composite sample shows significantly larger agglomeration of the nanotube
additives.

sample. The time-dependent strain response of the sample is the SWNT weight fraction is increased to 0.01%, a reduction
measured using an extensometer (MTS 632.26E-20) attached of about 13% in creep strain compared to the pure epoxy is
to the sample. First, a nano-composite sample with 0.1 wt% of observed. On further increasing the SWNT weight fraction to
SWNT additives was tested at a constant stress of 10 MPa and 0.1%, the reduction in creep strain improves to 30%. The creep
its creep response was compared to that of a pure epoxy sam- performance improvement then plateaus with increasing nan-
ple. The results are shown in figure 1(B); the nano-composite otube weight fraction (e.g. for 0.25% SWNT weight fraction,
exhibits a significant reduction in creep strain compared to the the creep reduction levels off at ∼27%). However, when the
pure epoxy. After 50 h of continuous testing, the creep strain SWNT weight fraction was raised to 1%, the creep strain after
for the nano-composite (∼0.9%) is over 30% lower than the 50 h of testing showed ∼90% degradation (figure 1(C)) com-
baseline epoxy material (∼1.3%). Next, samples with differ- pared to the pure epoxy. Therefore it appears that the desir-
ent weight fractions (0%, 0.001%, 0.01%, 0.1%, 0.25% and able SWNT weight fraction for minimum creep response is in
1%) were tested to study the effect of nanotube weight fraction the 0.1–0.25% range, with associated creep reduction of about
on the material creep response. Figure 1(C) shows the percent- 30% compared to the baseline. To investigate the reason for the
age reduction in creep strain for the nano-composite compared loss in creep performance for the 1 wt% SWNT sample, scan-
to the baseline epoxy sample after 50 h of continuous load- ning electron microscopy (SEM) characterization of the frac-
ing at a constant stress of 10 MPa. For a small weight frac- ture surface (figure 1(D)) of the nanotube–epoxy samples was
tion of SWNT additives (e.g. 0.001 wt%), no noticeable reduc- performed. The dispersion quality of SWNT in the 0.1 wt%
tion in creep strain over the pure epoxy is observed. When sample is quite uniform, as seen in the fracture surface image

2
Nanotechnology 18 (2007) 185703 W Zhang et al

A B

C D

Figure 2. Effect of applied stress and operating temperature on the creep response: (A) creep strain deformation for the 0.1 wt%
nano-composite and pure epoxy sample at an applied stress of 20 MPa and room temperature. (B) Corresponding results at an applied stress
of 30 MPa and room temperature. At the higher stress level the response of the 0.1 wt% nano-composite and pure epoxy samples are identical.
(C) Creep strain deformation for the 0.1 wt% nano-composite and pure epoxy sample at an operating temperature of 40 ◦ C. The applied stress
is 10 MPa for both samples. (D) Corresponding results at an operating temperature of 55 ◦ C. Temperature dramatically increases the creep
deformation of the baseline epoxy material but it does not significantly affect the ability of the nanotube additives to reduce the creep
response.

(top SEM images of figure 1(D)); similar results were also ob- creep strain reduction for the SWNT–composite after 50 h
served for the 0.25 wt% SWNT–epoxy sample. In contrast, is ∼30%. However, with an increase in stress, the creep-
for the 1 wt% SWNT–epoxy sample, there appears to be a sig- resistant behaviour of the SWNT composite slowly diminishes.
nificant agglomeration of SWNT, as observed in the bottom At 20 MPa stress, the creep strain reduction for the SWNT
SEM images of figure 1(D). These agglomerations decrease composite over the pure epoxy is only 17% (figure 2(A)). At
the number of nanotube–epoxy interfacial contacts and create a relatively larger stress level of 30 MPa, there is virtually
voids/defects in the material which could cause the creep rate no change (figure 2(B)) in the observed creep strain response
to increase, as observed in the experiments. Note that the static of the SWNT composite and the pristine epoxy material.
Young’s modulus for all the samples was measured prior to the
This result provides an important clue towards understanding
creep tests, and there was no significant difference in Young’s
the underlying mechanism for creep reduction in our system.
modulus between the samples with and without nanotube addi-
The ability of the nanotube additives to reduce creep appears
tives in the 0.001–0.25 wt% fraction range. However, the creep
to be strongly related to the nanotube matrix interfacial
modulus of the nano-composite showed significant difference
compared to the pure epoxy, as evidenced by the creep strain bonding. At the low stress levels, this interfacial bonding
data shown in figure 1. holds and the nanotube additives form an infiltrating network
Next, the effect of the material stress level on the creep of cross-linking elements that interconnect the epoxy chains,
deformation response was investigated. For this test the thereby increasing their effective cross-linking density and
SWNT–epoxy nano-composite with 0.1 wt% of nanotube restricting the localized re-orientation and disentanglement
additives was tested at three different stress levels (10, 20 of epoxy chains under mechanical stress. At the higher
and 30 MPa); the test results are shown in figures 1(B), 2(A) stress levels, the nanotube–polymer interfaces are prone to
and (B). At the relatively lower stress level of 10 MPa, the failure [15–19, 23–26] and, as a consequence, the ability of

3
Nanotechnology 18 (2007) 185703 W Zhang et al

A composite (figure 2(C)) is 28% lower than the pristine epoxy,


and a similar reduction can also be observed for the 55 ◦ C
test case (figure 2(D)). This indicates that the nanotube–matrix
interfaces are not weakened significantly by increasing the
temperature and, as a consequence, the ability of the nanotube
cross-linking elements to limit epoxy chain re-orientation is
not adversely affected by the temperature, which makes such
systems attractive for high-temperature structural applications.
The bump observed in figure 2(C) after ∼104 s of testing could
be caused by a temperature-dependent morphological change
or relaxation of the epoxy, which causes a sudden change in
the creep compliance. However, since this effect is felt equally
for both the pure epoxy sample and the nano-composite, the
comparison between the two is still valid.
B
Tests were also performed to compare the performance
of multi-walled carbon nanotubes (MWNT) and C60 fullerene
additives with that of the single-walled nanotube (SWNT)
results shown previously. In all cases, the exact same
fabrication protocols (figure 1(A)) were used and the weight
fraction of the additives was held constant at 0.1% for all the
samples. Figure 3(A) shows the creep strain time history at
an applied stress of 10 MPa, while figure 3(B) compares the
measured creep strain for the different samples after 3 h of
testing. The creep testing was limited to 3 h in this case, since
this time period is sufficient to compare the material creep
response of the various nano-composite samples. Clearly, the
SWNT additives outperform both the C60 and the MWNT
Figure 3. Effect of filler geometry on the creep response. (A) Creep additives. For the MWNT additives, only the outer shells are
strain deformation versus time for pure epoxy sample and epoxy in contact with the epoxy chains while the inner concentric
samples with 0.1 wt% of SWNT, MWNT and C60 additives. The cylinders play no role in increasing the effective cross-linking
tests are performed at room temperature and the applied stress is density within the nano-composite. Therefore, for the same
10 MPa. (B) Measured creep strain after 3 h of continuous testing for weight fraction of additives we expect that the SWNT fillers
the pure epoxy and nano-composite samples. For the same weight
fraction (0.1%), the SWNT additives offer significantly better creep will offer significantly higher resistance to the re-orientation of
resistance compared to the MWNT and the C60 additives. epoxy chains than the MWNT fillers, as was confirmed by our
results. Our results also indicate that, for 0.1% filler weight
fraction, SWNT additives provide a significant improvement
the nanotubes in providing a barrier for the polymer chain (figures 3(A) and (B)) in creep resistance compared to C60
movement is negated. additives. This effect may be related to the filler geometry;
The effect of operating temperature on the creep behaviour SWNT have significantly higher aspect ratios compared with
of the nano-composite was also investigated. For this test, spherical fullerene additives.
the 0.1 wt% SWNT–epoxy composite and the pure epoxy To summarize, we have shown that the addition of well-
sample were tested at three different operating temperatures dispersed single-walled carbon nanotube additives into an
(23, 40 and 55 ◦ C) and at a constant stress level of 10 MPa. epoxy host matrix significantly reduces the creep deformation;
Note that addition of as-received (i.e. non-functionalized) a reduction of up to 30% in the creep strain after only 50 h
nanotubes in low weight fractions (<0.5%) does not change of testing is demonstrated. Increasing the weight fraction
the glass transition temperature of the epoxy significantly, of nanotubes in the 0–0.1 wt% range improves the creep
as shown in several studies [28–30]. The test results after response; however, for larger weight fraction (∼1%) the
50 h of testing at 40 and 55 ◦ C are shown in figures 2(C) creep performance deteriorates rapidly due to poor-quality
and (D); corresponding results for the room-temperature case dispersion of the nanotubes in the epoxy resin. With an
(23 ◦ C) were shown previously in figure 1(B). The results increase in the applied stress, the creep resistance of the
indicate that the creep rate of the baseline epoxy increases SWNT additives is significantly reduced. We expect that
significantly at the elevated temperatures; for example, the this is due to the failure of the polymer–SWNT interface at
creep strain of the pure epoxy sample after 50 h of testing the high stress levels. However, increasing the temperature
at 10 MPa stress is ∼1.3% at room temperature, and this does not significantly impact the ability of the nanotube
increases to ∼4.75% at 40 ◦ C and ∼17.5% at 55 ◦ C. This additives to reduce the creep response of the nano-composite.
result is expected, since increased polymer chain mobility at SWNT additives also showed markedly higher creep strain
the elevated temperatures will facilitate the re-orientation and reduction compared to both MWNT and fullerene additives. In
disentanglement of the polymer chains, causing the creep to conclusion, carbon nanotube-reinforced polymer composites
increase. However, the ability of the SWNT additives to might play an important role in enabling the long-term higher-
reduce creep is not significantly diminished at the elevated temperature application of polymeric structures which would
temperatures. For example, at 40 ◦ C, the creep for the SWNT- otherwise fail through excessive creep deformation.

4
Nanotechnology 18 (2007) 185703 W Zhang et al

Acknowledgment [3] Bagchi A and Nomura S 2006 Compos. Sci. Technol. 66 1703
[4] Kymakis E and Amaratunga G A J 2004 Synth. Met. 142 161
This work was funded by the US National Science Foundation [5] Coleman J N, Curran S, Dalton A B, Davey A P, McCarthy B,
through the Faculty Early Career Development Award (CMS- Blau W and Barklie R C 1998 Phys. Rev. B 58 7492
0347604) to Nikhil Koratkar. [6] Coleman J N, Khan U, Blau W J and Gun’ko Y K 2006 Carbon
44 1624
[7] Zhang Q, Rastogi S, Chen D, Lippits D and Lemstra P J 2006
Appendix. Experimental details
Carbon 44 778
Figure 1(A) shows the procedure used to disperse the single- [8] Paiva M C, Zhou B, Fernando K A S, Lin Y, Kennedy J M and
Sun Y P 2004 Carbon 42 2849
walled carbon nanotube (SWNT) additives in the epoxy
[9] Allaoui W, Bai S, Cheng H M and Bai J B 2002 Compos. Sci.
system. The SWNT are purchased from CNI with a purity of Technol. 62 1993
over 95 wt%, length of 5–30 µm and diameter of 1–2 nm. The [10] Fisher F T, Bradshaw R D and Brinson L C 2002 Appl. Phys.
epoxy system is a two-part room-temperature-cured system Lett. 80 4647
with a glass transition temperature of 90 ◦ C. The required [11] Veedu V P, Cao A, Li X, Ma K, Soldano C, Kar S,
amount of SWNT (from 0.001 to 1.0 wt%) was first weighed Ajayan P M and Ghasemi-Nejhad M N 2006 Nat. Mater.
and mixed with 300 mg of the dispersion surfactant BYK-9076 5 457
[12] Zhang W, Suhr J and Koratkar N A 2006 Adv. Mater. 18 452
obtained from BYK Chemie Company. Next, 2.0 ml of acetone
[13] Chen W X, Li F, Han G, Xia J B, Wang L Y, Tu J P and
was added to the mixture; the mixture was stirred at 200 rpm Xu Z D 2003 Tribol. Lett. 15 275
for 15 min and was then placed into a room-temperature batch [14] Siegel R W, Chang S K, Ash B J, Stone J, Ajayan P M,
sonicator for 30 min. The mixture was added to epoxy (Epoxy Doremus R W and Schadler L S 2001 Scr. Mater. 44 2061
2000 from Fibreglast Inc.), stirred manually for 10 min and [15] Suhr J, Koratkar N, Keblinski P and Ajayan P M 2005 Nat.
then placed in the batch sonicator for a further 10 min. De- Mater. 4 134
gassing was then performed by heating the sample in a vacuum [16] Suhr J, Zhang W, Ajayan P M and Koratkar N A 2006 Nano
Lett. 6 219
oven at 50 ◦ C for 10 h; this step was essential to evaporate and
[17] Koratkar N A, Suhr J, Joshi A, Kane R S, Schadler L S,
remove the acetone in the mixture. The sample was then cooled Ajayan P M and Bartolucci S 2005 Appl. Phys. Lett.
and 75 mg (0.5 wt%) of an air-releasing agent (BYK A-500) 87 063102
was added to the mixture. The mixture was stirred manually for [18] Himanshu R and Nader J 2005 Compos. Sci. Technol. 65 2079
10 min and then de-gassed under vacuum at room temperature. [19] Zhou X, Shin E, Bakis C and Wang K W 2004 Compos. Sci.
Next, low-viscosity curing agent 2120 (from Fibreglast Inc.) Technol. 64 2425
was added and blended into the epoxy using a high-speed [20] Zhang Z, Yang J L and Friedrich K 2004 Polymer 45 3481
[21] Yang J L, Zhang Z, Schlarb A K and Friedrich K 2006 Polymer
mixer (Model ARE-250 from Thinky Corporation) at 2500 rpm
47 2791
for 4 min. The mixture was then placed in vacuum again for a [22] Vlasveld D P N, Bersee H E N and Picken S J 2005 Polymer
final de-gassing stage for 15 min and was poured slowly into 46 12539
silicone rubber molds. The epoxy composite was cured at room [23] Wagner H D and Vaia R A 2004 Mater. Today 38 205
temperature and a pressure of 90 psi for 24 h, followed by 2 h [24] Cooper C A, Cohen S R, Barber A H and Wagner H D 2002
of post-cure at 80 ◦ C. The high-pressure curing was necessary Appl. Phys. Lett. 81 3873
to prevent the formation of micro-bubbles in the composite [25] Yu M F, Yakobson B I and Ruoff R S 2000 J. Phys. Chem. B
104 8764
samples. Once the samples were cured, they were removed
[26] Barber A H, Cohen S R, Eitan A, Schadler L S and
from the mold and then polished using a 600-grid sandpaper Wagner H D 2006 Adv. Mater. 18 83
for a smooth surface finish. [27] Gojny F H, Wichmann M H G, Fiedler B and Schulte K 2005
Compos. Sci. Technol. 65 2300
References [28] Barrau S, Demont P, Maraval C, Bernes A and
Lacabanne C 2005 Macromol. Rapid Commun. 26 390
[1] Thostenson E T, Ren Z and Chou T W 2001 Compos. Sci. [29] Fidelus J D, Wiesel E, Gojny F H, Schulte K and Wagner
Technol. 61 1899 H D 2005 Composites A 36 1555
[2] Lau A K and Hui D 2002 Composites B 33 263 [30] Gojny F H and Schulte K 2004 Compos. Sci. Technol. 64 2303

You might also like