Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/234252195

Raman spectroscopy of carbon-nanotube-based composites: One


contribution of 13 to a Theme 'Raman spectroscopy in carbons: from
nanotubes to diamond'

Article  in  Philosophical Transactions of The Royal Society B Biological Sciences · October 2004


DOI: 10.1098/rsta.2004.1447 · Source: PubMed

CITATIONS READS

182 1,054

2 authors, including:

Daniel Hanoch Wagner


Weizmann Institute of Science
354 PUBLICATIONS   21,432 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Composite materials, interface adhesion View project

EU Matera grant „Bioactive Nanocomposite Constructs for Regeneration of Articular Cartilage.” Acronym:BioNanoCoRe (2007-2010y) View project

All content following this page was uploaded by Daniel Hanoch Wagner on 29 May 2014.

The user has requested enhancement of the downloaded file.


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of carbon−nanotube−based composites


Qing Zhao and H. Daniel Wagner
Phil. Trans. R. Soc. Lond. A 2004 362, 2407-2424
doi: 10.1098/rsta.2004.1447

Email alerting service Receive free email alerts when new articles cite this article - sign up in the box at the top
right-hand corner of the article or click here

To subscribe to Phil. Trans. R. Soc. Lond. A go to: http://rsta.royalsocietypublishing.org/subscriptions

This journal is © 2004 The Royal Society


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

10.1098/rsta.2004.1447

Raman spectroscopy of
carbon-nanotube-based composites
By Q i n g Z h a o a n d H. Daniel Wagner
Department of Materials and Interfaces, Weizmann Institute of Science,
Rehovot 76100, Israel (daniel.wagner@weizmann.ac.il)

Published online 20 September 2004

Recent developments in the application of Raman spectroscopy to carbon-nanotube-


based composite materials are reviewed. This technique may be used to identify
carbon nanotubes, access their dispersion in polymers, evaluate nanotube/matrix
interactions and detect polymer phase transitions. The Raman spectra of nanotubes
can also be used to quantify the strain or stress transferred to nanotubes from the
surrounding environment and to investigate local stresses and strains in polymers
and composites. A polarized Raman technique was developed to detect the stress or
strain in a matrix using randomly dispersed single-walled nanotubes. This technique
has been used to detect and map stress fields in model fibre–polymer composites. The
stress distributions around fibre breaks were mapped and compared with classical
load transfer models.
Keywords: Raman spectroscopy; carbon nanotube; composite

1. Carbon nanotubes: discovery, structure and applications


Carbon-based tubular nanostructures were reported in 1982 by Nesterenko et al ., but
their work remained largely unknown to the scientific community (Nesterenko et al.
1982). It is thanks to Iijima (1991), who reported the synthesis of carbon nanotubes
(CNTs) in 1991, following the discovery of fullerenes (Kroto et al. 1985), that over
the last decade CNTs have become an entirely new field of research.
The structure, topology and size of CNTs are the source of their outstanding
mechanical properties, fascinating electronic properties, and of a whole range of
promising applications (Dresselhaus 1995; Salvetat et al. 1999; Treacy et al. 1996;
Yakobson & Smalley 1997). The latter include their use as electron-field emitters for
vacuum microelectronic devices, nanoprobes at the tip of an atomic force microscope
(AFM), efficient supports in heterogeneous catalysis and a medium for lithium and
hydrogen storage. Nanotubes have been embedded into various materials to produce
composites with modified electrical conductivity, magnetic properties and optical
properties. Nanotubes are also promising candidates as a mechanical reinforcement
phase in composite materials.

One contribution of 13 to a Theme ‘Raman spectroscopy in carbons: from nanotubes to diamond’.

Phil. Trans. R. Soc. Lond. A (2004) 362, 2407–2424 


c 2004 The Royal Society
2407
Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2408 Q. Zhao and H. D. Wagner

1000
critical
900 strain
points
800
5% SWNT loading
700

stress (MPa)
600
500
400
300 1% SWNT loading
200
100
0
0 0.003 0.006 0.009 0.012 0.015
strain (mm mm−1)
Figure 1. Typical stress–strain plots for composite carbon fibres with 1 and 5 wt % SWNT
loadings. (Reproduced with permission from Andrews et al. (1999).)

2. Carbon nanotube composites and their properties


The sp2 carbon–carbon bond in the basal plane of graphite is the strongest of all
chemical bonds, but the weakness of the interplanar forces means that ordinary
graphite is of little value as a structural material. One way in which the great strength
of the sp2 bonds can be exploited practically is through a fibrous morphology in
which all basal planes run approximately parallel to the axis. Such fibres indeed exist
commercially but are not perfect: they contain various types of structural defects and
misaligned planes. The quasi-perfect structure of CNTs, however, opens a new route
to super-strong carbon fibres. Nonetheless, the potentially outstanding mechanical
properties of CNTs will be of little value unless nanotubes can be incorporated
into a matrix. At the same time, since nanotubes are electrically conductive (the
conductivity depends on their structure) and have greater diamagnetic susceptibility
than graphite, they are also capable of improving the electric or magnetic properties
of the polymer matrices (Harris 1999). For example, the presence of the nanotubes
in poly(vinyl alcohol) (PVOH) not only stiffens the material at high temperature
and retards the onset of thermal degradation, but it also leads to typical percolation
behaviour and to improved electrical conductivity compared with the PVOH matrix
(Shaffer & Windle 1999).
Mechanical tests show that extrusion produced single-walled nanotube/poly-
propylene (SWNT/PP) fibres with 1 wt % SWNTs have a tensile strength 40%
higher than pure PP. Moreover, the modulus of the composite fibres is 55% higher
than that of the PP fibre (Kearns & Shambaugh 2002). SWNT/pitch composite
fibres have been produced through a wind-up drum, and figure 1 shows the typical
stress–strain curves of composite fibres with 1 and 5 wt % nanotubes. The addition
of 5 wt % SWNTs increased the strength, modulus and conductivity of the composite
by ca. 90%, 150% and 340%, respectively (Andrews et al. 1999). In another work,
tensile tests of MWNT/polystyrene composites showed that the addition of 1 wt %
MWNTs increases the breaking stress and elastic modulus, and in situ transmission

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2409

electron microscopy (TEM) observation confirmed the effective load transfer between
the matrix and the nanotubes (Qian et al. 2000). In conclusion, nanotubes may in
principle be used as polymer modifiers to improve electric, magnetic and mechanical
properties.

3. Raman spectroscopy of CNTs


Various methods have been used to study the structure and physical properties of
CNTs. TEM and high-resolution transmission electron microscopy (HRTEM) are the
most powerful techniques for structural research. X-ray diffraction provides accurate
crystallographic measurements. Scanning tunnelling microscopy (STM) and atomic
force microscopy provide topological information with atomic resolution. Magnetic
force microscopy is used to characterize the dispersion of nanotubes in a polymer
matrix (Lillehei et al. 2002).
Raman spectroscopy is non-destructive and readily available, and measurements
can be made over a wide range of temperatures or pressures. It can provide unique
information about vibrational and electronic properties of the material. Even though
it is not a direct method, it can also be used to determine the structure of the material
and allows the identification of materials through the characteristic vibrations of
certain structures. Because the Raman intensity of a vibration or phonon in a crystal
depends on the relative directions of the crystal axis and the electric wave polarization
of the incident and scattered light, it may also be used to determine the orientation
of nanotubes in polymer matrices or within nanotube bundles.

(a) Raman spectra of CNTs


Raman spectroscopy has been used to determine the diameter of nanotubes, the
diameter distribution of nanotube bundles and the structural properties of nanotubes
(Dresselhaus et al. 2002a). The unique one-dimensional molecular nature of SWNTs
makes the resonance Raman technique an extremely useful and accurate method for
the identification of (n, m) indices, or alternatively of the diameter and the chirality
of an individual SWNT. By knowing the (n, m) vector, the dependence of all the
features of the spectra on the diameter, chiral angle, laser excitation energy and
other parameters can be worked out in detail. Therefore, the spectrum of SWNT
bundles can be interpreted and the effect of nanotube–nanotube interactions can be
deduced (Dresselhaus et al. 2002b).
Figure 2 shows the Raman spectrum of a single SWNT (Dresselhaus et al. 2002b).
A prominent feature is the radial breathing mode in the 160–300 cm−1 region, asso-
ciated with a symmetric movement of all carbon atoms in the radial direction. Both
theoretical and experimental work has shown that the frequency of the radial breath-
ing mode is inversely proportional to the diameter of individual nanotubes. In nano-
tube aggregates, it has been postulated that there is a weak inter-tubule coupling
via van der Waals interactions that manifests a 6–20 cm−1 shift to higher frequencies
due to the space restrictions imposed by neighbouring tubes (Venkateswaran et al.
1999). The D Raman band of CNTs, which is observed between 1250 and 1450 cm−1 ,
has a linear dependence on the laser excitation energy. This band is activated in the
first-order scattering process of sp2 carbons by the presence of in-plane substitu-
tional hetero-atoms, vacancies, grain boundaries or other defects and by finite-size

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2410 Q. Zhao and H. D. Wagner

(a) (b)

Raman intensity

9
RBM D G'

number
G 6
3
2nd 0 1 2 3
0.5 µm
dt (nm)
0 1000 2000 3000
frequency (cm−1)
Figure 2. (a) Raman spectrum from one nanotube taken over a broad frequency range using
Elaser = 785 nm = 1.58 eV excitation, and showing the radial breathing mode, the D band, the
G band, second-order features and the G band. The features marked with ‘∗’ at 303, 521 and
963 cm−1 are from the Si/SiO2 substrate and are used for calibration of the nanotube Raman
spectrum. (b) AFM image of the sample showing isolated SWNTs grown from the vapour phase.
The small particles are iron catalyst particles. The inset shows the diameter distribution of
this sample based on the AFM observations of 40 SWNTs; dAV t = 1.85 nm. (Reproduced with
permission from Dresselhaus et al. (2002a).)

effects, all of which lower the crystalline symmetry of the quasi-infinite lattice. This
association of the D band with symmetry-breaking phenomena results in a D band
intensity that is proportional to the phonon density of states (Brown et al. 2001).
The G (or D∗ ) band is the second-order overtone of D band. The locations of D
and D∗ Raman bands of CNTs depend linearly on the laser excitation energy and
this dispersion relation of the D and D∗ band of nanotubes is similar to that in other
sp2 carbons, though with some distinct characteristic behaviour that is specific to
nanotubes. A shift of the D and D∗ bands of CNTs has been observed in different
cases. The strong linear dispersion of the D and D∗ band frequency in graphite
has been identified with a resonance process between a phonon and an electronic
transition between π and π ∗ electronic states (Matthews et al. 1999; Pócsik et al.
1998; Thomsen & Reich 2000).

(b) Existence and states of CNTs by Raman spectroscopy


Since CNTs have distinctive Raman spectra, Raman spectroscopy can be used to
prove the existence of nanotubes. For instance, the Raman spectra of the carbonized
products showed that poly(p-phenylene vinylene) (PPV) nanotubes and nanorods
are converted into corresponding CNTs and nanorods after 1 h heat treatment at
850 ◦ C (Kim & Jin 2001). The fact that the RBM vibrational frequencies are sensi-
tive to the nanotube diameter and that the intensity of the spectra is proportional
to the number of nanotubes makes it possible that Raman spectroscopy can pro-
vide even more detailed information. For example, capillary electrophoresis (CE) has
been performed on polymer-stabilized bundles and sodium dodecyl sulphate (SDS)
suspensions of SWNTs. Real-time Raman spectra of samples with three individual

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2411

nanotube fractions were compared and the Raman frequencies were identical for
all three fractions. This indicates that each fraction contained nanotubes of equiv-
alent diameter. There was no obvious band broadening and shoulders observed in
the sample, which suggests a high degree of uniformity in nanotube diameter within
the sample. Therefore, the difference in mobility between fractions is due to differ-
ences in nanotube lengths (Doorn et al. 2002). More results indicate that CE has the
potential to separate CNTs, not only by differences in length but also by differences
in size or other geometric factors, such as diameter or cross-section. Closer inspec-
tion of how the radial breathing mode region of the Raman spectra changes over
the course of a separation, provides information on whether diameter-selective sepa-
rations are occurring superimposed on the length-based mechanism. The variations
in the radial breathing mode intensities indicate that there are diameter-dependent
spectral changes occurring over the course of a separation (Doorn et al. 2003).
The Raman frequency, intensity and band shape of the vibrations can vary when
nanotubes interact with other species: this can be used to examine the structure
of the interface and obtain information about the nature, localization and force
of the interaction. Raman spectroscopy was used to probe the structure of pep-
tide/nanotube fibres, in which a specially designed amphiphilic α-helical peptide
was used not only to coat and solubilize CNTs, but also to control the assembly
of the peptide-coated nanotubes into macromolecular structures through peptide–
peptide interactions between adjacent peptide-wrapped nanotubes. After the nano-
tubes were dispersed in aqueous solution by sonication, a shift of more than 9 cm−1
was observed towards higher frequencies in the radial breathing mode, evidence that
nanotubes were no longer in direct contact with one another along their length as
they were in the absence of peptide. The changes observed in the nanotube tangential
mode feature centred at 1590 cm−1 emphasized that peptide is coating the individual
nanotubes in the peptide/SWNT composite. Raman spectra of NaCl induced fibres
confirmed the presence of CNTs in the fibres and showed increases in the radial
breathing mode frequencies, suggesting that the fibres were composed of peptide-
coated SWNTs (Dieckmann et al. 2003).

4. Raman spectroscopy of CNT composites


In this section we demonstrate that CNTs embedded in polymers may achieve several
roles, including use as orientation detectors as well as molecular sensors around
structural defects and reinforcements. Previously unpublished data are presented
about related temperature effects.

(a) Orientation of nanotubes in polymers


SWNTs show resonance-enhanced Raman scattering effects when a visible or near
infrared laser is used as the excitation source, whereas most polymers do not show
such a resonance effect. Since in such case a large difference in intensity exists between
the SWNT peaks and those arising from the polymer, Raman spectroscopy is an ideal
characterization technique for the orientation study of SWNT. Indeed it was observed
that the Raman spectra intensity of the tangential mode (1500–1700 cm−1 ) monoton-
ically decreases with increasing the angle between the fibre axis, and the polarization
direction of the polarizer has been used to show the orientation effect. For example,

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2412 Q. Zhao and H. D. Wagner

polarized Raman spectroscopy has been used to analyse the orientation of CNT
bundles generated by an electrophoretic method (Poulin et al. 2002). In another
case, nanotubes were oriented into bundles by applying an electric field between a
carbon fibre and the ultrasonicated SWNT/N, N-dimethylformamide (DMF) suspen-
sion, and polarized Raman spectroscopy was then used to quantify the alignment of
nanotubes as a function of the angle between the fibre and the polarizer (Gommans
et al. 2000). The crystallite orientation and the SWNT alignment in melt-blended
SWNT/PP composites fibres have been studied using X-ray diffraction and polarized
Raman spectroscopy, which showed that there is an orientation effect in the drawn
SWNT/PP fibre (Bhattacharyya et al. 2003). A combination of solvent casting and
melt mixing was used to disperse SWNTs in poly(methyl methacrylate) (PMMA),
and polarized Raman spectroscopy was again used to demonstrate the alignment of
nanotubes in the PMMA (Haggenmueller et al. 2000). More recently, Frogley et al.
(2002) have performed a thorough study of nanotube alignment in polymers using
polarized Raman spectroscopy, and compared a large amount of experimental results
with existing models such as those by Saito et al. (1998) and Gommans et al. (2000).
It was demonstrated that a non-resonant theory (Saito’s) gives the most reasonable
orientation distribution for the tubes. A good degree of nanotube alignment was
obtained by a simple shear-flow technique.

(b) Nanotube–polymer interactions


Raman spectroscopy has been used to probe the interaction between polymers and
nanotubes in nanotube-based composites. Generally, the interaction between nano-
tubes and polymer is reflected by a peak shift or a peak width change. This is the
case in composites made of SWNT-filled polystyrene (PS) and styrene isoprene (SI),
prepared by a mini-emulsion polymerization process (Barraza et al. 2002). No visible
change in the polymer peak locations as a result of the insertion of nanotubes could
be detected. However, the RBM of the SWNTs was observed to shift to lower frequen-
cies, and a split into two low-intensity broad peaks of the initially sharp peak was
also observed. These results from the breathing mode of SWNTs are largely affected
by the surroundings of the individual tube. For instance, the outer diameter of the
nanotubes is reduced as a consequence of the physical constraint introduced by the
surrounding polymer chains, and this radial deformation is believed to contribute
to a local mechanical interlocking mechanism for nanotubes. In another example,
the nanotube Raman breathing modes shift upwards following an acid treatment,
because acid molecules are intercalated in the nanotube ropes, though the original
frequencies are recovered after high-temperature annealing. In the same manner, the
shifts observed with the introduction of nanotubes in PMMA can be explained by the
intercalation of polymer into bundles. In fact, the polymer exerts a pressure on the
individual tubes, thereby increasing the breathing mode frequencies (Rinzler et al.
1998; Stephan et al. 2000). In the case of SWNT/PP matrix composites, the charac-
terization of these new materials was performed by differential scanning calorimetry
and Raman spectroscopy to obtain information on the crystallization kinetics of
PP and the macrostructure and organization of the nanotubes in the composites
(Stephan et al. 2000). The incorporation of SWNTs accelerates the nucleation and
growth mechanisms of PP crystals, this being more evident when the nanotube con-
tent is lower, because of bundle effects at higher nanotube contents (Valentini et al.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2413

2003). With the introduction of nanotubes in PP, the feature at 1270 cm−1 is shifted
up to 1275 cm−1 and its width is narrower. The peak observed at 1260 cm−1 for neat
SWNTs decreases in intensity in the composite, indicating a dilution effect of the
CNTs when blended with PP. In fact, Valentini et al. (2003) believe that the polymer
exerts a pressure on the individual tubes, which leads to an increase of the breathing
mode frequencies. In a recent study (Zhao et al. 2001b), irregularities (or discontinu-
ities) have been observed in the wavenumber–temperature plot for both amorphous
bisphenol A polycarbonate (PC) and polyurethane acrylate (PUA), where nano-
tubes have been embedded. The Raman results have been compared with dynamic
mechanical thermal analysis (DMTA) data for both polymers, and the sources of the
discontinuities were investigated. The concordance found between the DMTA data
and the Raman spectral response shows that these irregularities reflect basic poly-
mer phase transitions, which may thus be sensed by nanotubes and reflected in the
Raman spectrum. This demonstrates that SWNTs can be used as molecular sensors
(Zhao et al. 2001b).
These examples show that, since nanotubes are sensitive to their environment,
Raman spectroscopy is a useful and reliable tool for the investigation of nanotube–
polymer interactions and can be used for the research of CNT-based composites, as
will now be further demonstrated.

(c) Sensitivity of CNTs to stress and strain



D Raman wavenumber shifts are observed when SWNTs are dispersed in liquids.
The extent of shifting compared with the initial D∗ wavenumber in air depends on
the nature of the liquid. A correlation was found between the cohesive energy density
(CED, or, more loosely, the surface tension) of a medium and the D∗ wavenumber
(Wood et al. 2000). To corroborate this, further research was done by using a diamond
anvil cell (DAC) to apply hydrostatic stress while the Raman response of both the
G and D∗ peaks was monitored in situ. A striking similarity was indeed observed
between the hydrostatic pressure experiments, on the one hand, and the molecular
pressure experiments with different media, on the other hand, regarding the degree
of Raman wavenumber shift (Wood et al. 1999, 2000). In figure 3, the squares show
the D∗ Raman wavenumber of SWNTs in various media (having different CEDs),
whereas the triangles show the dependence of the D∗ band on applied hydrostatic
pressure. Both sets of data agree with each other, which demonstrates the sensitivity
of nanotubes to molecular pressure from the surrounding medium, as measured either
by CED or by mechanical pressure using a DAC.
The D∗ Raman wavenumber shift can also be related to the mechanical deforma-
tion of nanotubes. In the field of fibre composite materials, it has been known for
about two decades that the application of a mechanical strain to fibres such as carbon
or Kevlar results in shifted frequencies of the Raman bands which are directly related
to the inter-atomic force constants. Correlating such shifts with the applied strain
(through a calibration procedure) leads to the determination of local stress profiles
in the embedded fibres. A similar effect is observed in CNT composites. The tensile
and compressive properties of MWNT/epoxy composites and pure epoxy were mea-
sured and the D∗ Raman spectra obtained (Schadler et al. 1998). The D∗ peak was
observed to shift upwards by 7 cm−1 under a 1% compressive strain. A slightly pos-
itive shift appears when the composites are under tension. The shifts are thought to

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2414 Q. Zhao and H. D. Wagner

arise from the strain transferred from the matrix to nanotubes. The different Raman
responses in tension and compression are most likely to be due to the structure of
MWNTs. Indeed, Schadler et al. (1998) believe that, under tension, the outer layer
of the MWNT is loaded, but load is not effectively transferred to the inner layers
due to the relatively weak bonding between the nanotube layers. Instead, the inner
tubes may slip with respect to the outer tubes. Since the Raman signal is averaged
over the whole MWNT, the result is only an insignificant Raman peak shift. Under
compression, however, the load transfer to the inner layers of the MWNT occurs
through buckling and the bent sections of the nanotubes. Slippage of nanotubes lay-
ers in compression is prevented because of the seamless structure of the tubes and
the geometrical constraint the outer layer imposes on the inner layers (Schadler et al.
1998). Similar experiments have been performed with SWNTs, but the D∗ Raman
results showed that under compression there is almost no shift and only a small
downward shift under tension (Ajayan et al. 2000). Ajayan et al. (2000) observed
that the applied compressive stress is transferred into buckling, bending or twisting
of the nanotube network without introducing important local deformation that can
be monitored by Raman. Since SWNTs tend to form ropes, the small Raman shift
under tension is due to the fact that the individual SWNTs are slipping within the
ropes and decrease the load required to deform the ropes (Ajayan et al. 2000). In
both cases, the tensile and compressive strains are applied uniaxially on the samples,
though the nanotubes are dispersed randomly in the polymer matrix (Ajayan et al.
2000; Schadler et al. 1998). When the nanotube composite samples are under ten-
sion, nanotubes lying along the tensile direction are under tension, but those in the
perpendicular direction are under compression because of the Poisson’s contraction.
The reverse is true when the sample is under compression. Since the Raman laser
spot size is ca. 1–5 µm thick, and nanotube diameters are several nanometres in size,
the Raman signal is averaged over nanotubes in all directions, but Ajayan et al.
(2000) and Schadler et al. (1998) do not consider this in detail.
Significant D∗ Raman shifts were measured when a nanotube polymer composite
was tested at different temperatures under no mechanical tension, and this was con-
sidered to result from thermal shrinkage (Lourie & Wagner 1998). Other results show
that an elastic tensile strain can be transferred from the polymer matrix to SWNTs
resulting in a downward shift of the D∗ Raman wavenumber of the nanotubes (Cooper
et al. 2001). Additional research shows that an empirical linear relationship exists
between the SWNT D∗ wavenumber shift and the applied elastic strain (Wood et al.
2001). If the nanotube D∗ wavenumber difference between zero strain and the applied
strain (ε) is defined as the Raman wavenumber shift ∆Nw , then the empirical slope
m of the wavenumber–strain relation can be defined from:
∆Nw = mε. (4.1)
A mechanical stress–strain curve can be recorded simultaneously with the Raman
measurement. In the elastic regime of a tensile test, the stress σ and the strain ε in
the polymer are related by
σ = Eε, (4.2)
where E is the Young modulus of the polymer. Therefore, the matrix stress may be
deduced from the Raman measurement:
E∆Nw
σ = Eε = . (4.3)
m
Phil. Trans. R. Soc. Lond. A (2004)
Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2415

2635
10 13 14
2630 9 15
wavenumber (cm−1) 6 12 16
11
2625 5
7
3
2620 2 8
4
2615

2610 1

2605
0 400 800 1200 1600 2000 2400 2800
pressure (MPa)

Figure 3. The peak positions of the D Raman peak in SWNTs plotted as a function of () the
molecular (by immersion in liquids) and () macroscopic (using DAC) pressure. The molecular
pressure is the cohesive energy density of the liquid in which the nanotubes are immersed: for
these experiments the external applied pressure was ambient, 1 atm. The macroscopic pressure
was applied to nanotubes immersed in liquid argon as a hydrostatic pressure medium. The solid
line is a polynomial fit. 1, air; 2, decane; 3, hexane; 4, dodecane; 5, cyclohexane; 6, carbon
tetrachloride; 7, chloroform; 8, hexylene glycol; 9, acetone; 10, diethylene glycol; 11, propylene
glycol; 12, ethanol; 13, ethylene glycol; 14, glycerol; 15, formamide; 16, water. (Reproduced with
permission from Wood et al. (2000).)

2645

2640
m (235 K) = 909 cm−1 / ε
wavenumber (cm−1)

2635

m (298 K) = 467 cm−1 / ε


2630

2625

2620

2615
0 0.01 0.02 0.03 0.04 0.05
strain
Figure 4. The wavenumber–strain response of SWNTs embedded in PUA.  data are obtained
at 298 K (room temperature) and • data are obtained at 235 K.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2416 Q. Zhao and H. D. Wagner

140

120

100

σy
stress (MPa)

80

60

40 σy

20

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07


strain
Figure 5. Mechanical and spectroscopic stress–strain curves, as determined from the Raman
shift data in figure 4.  data are obtained at 298 K (room temperature) and • data are obtained
at 235 K. σy is the yield point of the polymer according to the mechanical stress–strain curve.

Experiments show that m varies when nanotubes are embedded in different poly-
mer matrices (Frogley et al. 2002; Wood et al. 2001; Zhao et al. 2001b). Recent results
in our laboratory show that, for a given matrix, m is a temperature-dependent param-
eter. Figure 4 shows the Raman wavenumber–strain response of SWNTs embedded
in PUA at two temperatures. The data denoted by triangles were obtained at 298 K
(room temperature) and the data denoted by solid circles were obtained at 235 K.
The initial part of both datasets is approximately linear within the elastic strain
region (ca. 1.5%). The values of m at both temperatures are 467 cm−1 /ε (as in Wood
et al. (2001)) and 909 cm−1 /ε, respectively. As the tensile strain increases, the wave-
number reaches a value of ca. 2623 cm−1 then remains approximately constant with
increasing strain, for both temperatures. The temperature dependence of the m val-
ues of the SWNTs in PUA can be attributed to the fact that the Young modulus (E)
of the polymer is a temperature-dependent parameter, E(T ). Equation (4.3) should
read
E(T )∆Nw
σ = E(T )ε = . (4.4)
m(T )
A spectroscopic stress–strain curve can thus be constructed from these data. In fig-
ure 5, the solid lines represent the mechanical stress–strain curves of PUA at 235 K
and 298 K. The symbols in figure 5 are the ‘spectroscopic stress–strain curve’ con-
structed from equation (4.4) using the Raman data in figure 4. The values of m(T ),
E(T ) used for these specimens are obtained experimentally at the corresponding
temperatures. Within the linear region, the spectroscopic curves fit exactly the cor-
responding mechanical curves. Beyond ca. 1.5%, the spectroscopic data deviate from

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2417

3.5

3.0

normalized stress (σyy / σ 0) 2.5

2.0

1.5

1.0
theory
0.5

0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5


number of radii from hole centre
Figure 6. The normalized stress along the x-axis from the edge of a circular hole, based on the
D∗ peak shift of SWNTs in PUA. Applied loads (σ0 ) of () 4 MPa, (◦) 6 MPa and () 8 MPa
were employed. The solid line is the linear elastic solution. (Reproduced with permission from
Zhao et al. (2001c).)

the mechanical curve. The yield stress σy at which the mechanical stress–strain curve
becomes nonlinear is indicated on the figure.
The discussion thus far shows that SWNTs indeed perform as sensors to detect
the elastic stress or strain in the polymer matrix. Such a microscale Raman sensing
technique has practical importance. As an example, it may be useful to measure the
matrix stress distribution in the vicinity of fibres to detect or predict the onset of fail-
ure in composite materials (Hull 1981). As already mentioned, Raman spectroscopy
has been successfully used since the early 1980s to monitor the deformation of spe-
cific fibres such as aramid, SiC and carbon, as specific Raman peak frequencies of the
Raman bands of these fibres are indeed sensitive to the applied strain (Batchelder
& Bloor 1979; Galiotis et al. 1983; Robinson et al. 1987a). Correlating these Raman
shifts to the applied strain leads to an evaluation of the stress distributions in the
fibres at a micrometre scale, and provides a means of deducing the interfacial shear
stress (Galiotis & Batchelder 1988; Galiotis et al. 1984; Nielsen & Pyrz 1999; Robin-
son et al. 1987b; Wagner et al. 2000). In reality, the micro-Raman technique cannot be
applied universally, since some fibres do not have strain-sensitive Raman peaks, glass
fibres being the best known example. In addition, since most polymers do not have
strain-sensitive Raman bands, it is practically impossible to use Raman spectroscopy
in those materials to detect the stress (or strain) distribution around discontinuities
in polymers for most composites; it is also impossible to perform in situ measure-
ments of the stress (or strain) distribution in glass fibres and in the matrix. This is
where dispersing SWNT sensors in the polymer matrix proves useful, as a flexible
tool that makes such measurements possible.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2418 Q. Zhao and H. D. Wagner

(d ) Stress distribution around a discontinuity


To demonstrate that embedded nanotubes can measure stress fields, a classical
elasticity problem with a known analytical solution was selected, namely, a circular
hole in a plate under uniaxial tension (Dally & Riley 1985). A first issue that needed
to be addressed concerned the orientation of the nanotubes within the (polymer)
plate. When the tubes are randomly oriented in the plate, in the case of a uniaxial
stress, as in simple mechanical tension, Poisson contraction occurs in the transverse
direction. Thus, some nanotubes will be under compression while others will be under
tension, resulting in a mixed signal. Aligning the nanotubes is thus preferable in order
to measure the specific components of the complex strain or stress distributions. A
simple shear-flow method can be developed to orient nanotubes in a polymer such
as PUA, to certain degree. The elastic strain dependence of the D∗ wavenumber
shift was indeed measured with the loading parallel and perpendicular to the flow
direction. A significant difference between the two cases was found, indicating that
this orientation method is effective (Wood et al. 2001). Based on the fact that SWNTs
can be oriented in PUA, the associated stress components can be separated and the
stress distribution in the plate around the circular hole under uniaxial tension could
be mapped. This is shown in figure 6. The experimental results fit the linear elastic
solutions very well, showing that SWNT sensors are able to give quantitative stress-
field-distribution information at a stress discontinuity (Zhao et al. 2001c).

(e) Polarized Raman spectroscopy for stress mapping


in model fibre–polymer composite
In the example just presented, oriented nanotubes in a specific direction were nec-
essary to allow separation of the stress components, using a sample under mechan-
ical stress. This requirement restricts the application of the method because it is
not always straightforward to induce shear flow and orient nanotubes in a polymer.
An alternative is to use plane-polarized Raman measurements. In this case, the mea-
sured D∗ Raman peak intensity from a single nanotube is high when the polarization
direction is parallel to the nanotube length, and low when the polarization direction
is perpendicular to the nanotube length, as predicted by Saito et al. (1998). This
means that randomly oriented SWNTs in a polymer can act as strain sensors using
polarized Raman spectroscopy, rather than using specimens with oriented nanotubes
in the polymer, provided that a sufficiently high density of nanotubes is present in
the polarization direction.
Polarized Raman experiments under uniaxial strain have been performed with
nanotube composites, to demonstrate that the measured Raman wavenumber shift
with strain is indeed dependent on the optical polarization direction. The D∗ peak
position of SWNTs embedded in PUA was measured as a function of tensile strain, for
both uniaxial and random tube orientations. The results showed that the polarized
Raman can ‘select’ those nanotubes that are lying in one direction (Frogley et al.
2002). The orientation effect of the shear-flow method was estimated quantitatively.
The advantage of this technique is that, when nanotubes are used as strain sensors,
randomly oriented nanotubes combined with polarized Raman may be used to detect
the stress or strain distribution in any direction in the polymer matrix.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2419

1.00
4 1.06
R (fibre radius)

1.11
1.14
2 1.18

1.24

1.40 1.38 1.32


0
0 2 4 6
X (fibre radius)
Figure 7. Two-dimensional contour map for the stress concentration factor (Kc ) in the vicinity
of a fibre break in E-glass. The break is located at the origin and only a quarter of the map
is shown for simplicity. The X-axis runs along the fibre length and the Y -axis is perpendicular
to the fibre length. The unit of length on both axes is the fibre radius. The highest stress
concentration occurs just near the fibre break (Kc = 1.42) and is indicated in red. (Reproduced
with permission from Zhao & Wagner (2003).)

Further convincing evidence for the exploitation of nanotubes as mechanical sen-


sors with polarized Raman was demonstrated in complex stress state systems, by
monitoring the stress in the matrix around a fibre end (Zhao et al. 2001a) and a
single fibre break (Zhao & Wagner 2003). In composite material research, a single
fibre composite is often used to quantify the interfacial adhesion, as the properties of
the composite depend not only on the properties of the fibre and the matrix, but also
on the quality of the interface between them. The single fibre fragmentation tech-
nique is a known, practically convenient and reproducible method to characterize
the fibre–matrix adhesion or interfacial properties. However, it also is a complex test
which also involves shear yielding of the matrix, interfacial debonding and trans-
verse matrix cracking (Kelly & Tyson 1965). The occurrence of these additional
damage events during the fragmentation test makes the conventional data-reduction
technique (based on the constant-shear model) problematic. Thus, a determination
of the stress distribution around the fibre break in the matrix is necessary. Two-
dimensional stress profiles around a glass fibre break were mapped using SWNT
sensors randomly dispersed in a PUA matrix, by means of polarized Raman spec-
troscopy. A contour map showing the distribution of the stress concentration factor
around a fibre break was produced, as shown in figure 7. The stress concentration
effect decreases in both directions with distance away from the break point (Zhao &
Wagner 2003).
A final experiment combining the Raman sensitivity both of a carbon fibre and
of CNTs dispersed in the polymer around the fibre was performed as follows. The

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2420 Q. Zhao and H. D. Wagner

(a)
2.5

2.0

stress in the fibre (GPa)


1.5

1.0

0.5

−0.5

21 (b)
stress in the matrix (MPa)

18

15

12

6
−1000 0 1000 2000 3000
distance along the fibre (µm)
Figure 8. Stress distributions in (a) the HMCF and (b) the PUA matrix along the fibre edge
(2 µm from the edge), measured simultaneously by micro Raman spectroscopy. The distributions
are mirror images of each other. The applied stress level was 10 MPa. The solid line in (a) is a
polynomial fit to the data. In (b), the line is the mirror image of the fit in (a), scaled accordingly.
(Reproduced with permission from Zhao & Wagner (2003).)

basic idea of this experiment was to simultaneously detect the stress distributions
in the fibre and in the matrix using Raman spectroscopy, and to compare the
results with existing stress-transfer models. A continuous high-modulus carbon fibre
(HMCF)/PUA composite with nanotubes dispersed in the matrix was chosen. The
G Raman band of HMCF shifts linearly with applied strain. The strain in the fibre
and the matrix were measured simultaneously under three applied stress levels, par-
allel to the fibre direction. Figure 8 presents the distribution of stress in the carbon
fibre and in the matrix, between two break locations measured at an applied stress
of 10 MPa. A polynomial curve was fitted through the experimental data points. As
seen, the stress profiles in the matrix and the fibre, measured simultaneously for the
first time to the authors’ knowledge, complement each other well (Young et al. 2001).
Also, the experimental result fits the Kelly–Tyson stress-transfer analysis (Kelly &
Tyson 1965).

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2421

5. Conclusions
Recent developments in the application of Raman spectroscopy to CNT-based com-
posite materials have been reviewed, and some new data presented. This technique
may be used to distinguish CNTs from other carbon materials or polymers, as well
as to sort nanotubes by diameter or length. Raman spectroscopy has also been used
to check the dispersion of nanotube in polymers, evaluate nanotube/matrix inter-
actions, and detect polymer phase transitions. The Raman spectra of nanotubes
can also be used to quantify the strain or stress transferred to nanotubes from the
surrounding environment. Since small amounts of embedded nanotubes required to
make the polymer Raman sensitive to strain do not affect the mechanical properties
of the embedding matrix, these nanoscale tubes may be used to investigate local
stresses and strains in polymer materials, a result that is so far inaccessible by using
other methods. Polarized Raman can be used to detect the orientation of nanotubes
in polymer matrices. Oriented SWNT sensors can be used to detect stress fields
around a circular hole in a plate, demonstrating the applicability of the technique
for the mapping of the stress distribution in the vicinity of discontinuities.
A polarized Raman technique was developed to detect the stress or strain in a
matrix using randomly dispersed SWNTs, based on the fact that the intensity of the
signal arising from SWNTs is a function of the polarization direction. This method is
more convenient than the tube orientation technique, since it is not always possible
to orient SWNTs inside a polymer. The polarized Raman technique was used to
detect and map the stress fields in the model fibre–polymer composite. The stress
distributions around fibre breaks were measured and compared with classical load
transfer models in composite materials.
This project was supported by the (CNT) Thematic European network on ‘Carbon Nanotubes
for Future Industrial Composites’ (EU), the G. M. J. Schmidt Minerva Centre of Supramolecular
Architectures, and by the Israeli Academy of Science. H.D.W. is the incumbent of the Livio Norzi
Professorial Chair. Thanks are due to Dr Mark Frogley and Dr Jonathan Wood for guidance.
H.D.W. acknowledges the inspiring assistance of B. Goodman and L. Hampton.

References
Ajayan, P. M., Schadler, L. S., Giannaris, C. & Rubio, A. 2000 Single-walled carbon nanotube–
polymer composites: strength and weakness. Adv. Mater. 12, 750–753.
Andrews, R., Jacques, D., Rao, A. M., Rantell, T., Derbyshire, F., Chen, Y., Chen, J. & Haddon,
R. C. 1999 Nanotube composite carbon fibers. Appl. Phys. Lett. 75, 1329–1331.
Barraza, H. J., Pompeo, F., O’Rear, E. A. & Resasco, D. E. 2002 SWNT-filled thermoplastic
and elastomeric composites prepared by mini-emulsion polymerization. Nano Lett. 2, 787.
Batchelder, D. N. & Bloor, D. 1979 Strain dependence of the vibrational-modes of a diacetylene
crystal. J. Polym. Sci. B 17, 569–581.
Bhattacharyya, A. R., Sreekkumar, T. V., Liu, T., Kumar, S., Erincson, L. M., Hauge, R. H.
& Smalley, R. E. 2003 Crystallization and orientation studies in polypropylene/single wall
carbon nanotube composite. Polymer 44, 2373–2377.
Brown, S. D. M., Jorio, A., Dresselhaus, M. S. & Dresselhaus, G. 2001 Observations of the
D-band features in the Raman spectra of carbon nanotubes. Phys. Rev. B 64, 073403.
Cooper, C. A., Young, R. J. & Halsall, M. 2001 Investigation into the deformation of carbon
nanotubes and their composites through the use of Raman spectroscopy. Composites A 32,
401–411.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2422 Q. Zhao and H. D. Wagner

Dally, J. W. & Riley, W. F. 1985 Experimental stress analysis. McGraw-Hill.


Dieckmann, G. R., Dalton, A. B., Johnson, P. A., Razal, J., Chen, J., Giordano, G. M., Munoz,
E., Musselman, I. H., Baughman, R. H. & Draper, R. K. 2003 Controlled assembly of carbon
nanotubes by designed amphiphilic peptide helices. J. Am. Chem. Soc. 125, 1770–1777.
Doorn, S. K., Fields, R. E., Hu, H., Hamon, M. A., Haddon, R. C., Selegue, J. P. & Majidi, V.
2002 High-resolution capillary electrophoresis of carbon nanotubes. J. Am. Chem. Soc. 124,
3169–3174.
Doorn, S. K.. Strano, M. S., O’Connell, M. J., Haroz, E. H., Rialon, K. L., Hauge, R. H. &
Smalley, R. E. 2003 Capillary electrophoresis separations of bundled and individual carbon
nanotubes. J. Phys. Chem. B 107, 6063–6069.
Dresselhaus, M. S., Dresselhaus, G. & Saito, R. 1995 Physics of carbon nanotubes. Carbon 33,
883–891.
Dresselhaus, M. S., Jorio, A., Souza Filho, A. G., Dresselhaus, G. & Saito, R. 2002a Raman
spectroscopy on one isolated carbon nanotube. Physica B 323, 15–20.
Dresselhaus, M. S., Dresselhaus, G., Jorio, A., Souza Filho, A. G. & Saito, R. 2002b Raman
spectroscopy on isolated single wall carbon nanotubes. Carbon 40, 2043–2061.
Frogley, M. D., Zhao, Q. & Wagner, H. D. 2002 Polarized resonance Raman spectroscopy of
single-wall carbon nanotubes within a polymer under strain. Phys. Rev. B 65, 113413.
Galiotis, C. & Batchelder, D. N. 1988 Strain dependences of the 1st-order and 2nd-order Raman
spectra of carbon fibers. J. Mater. Sci. Lett. 7, 545–547.
Galiotis, C., Young, R. J. & Batchelder, D. N. 1983 A resonance Raman-spectroscopic study of
the strength of the bonding between an Epoxy-resin and a polydiacetylene fiber. J. Mater.
Sci. Lett. 2, 263–266.
Galiotis, C., Young, R. J., Yeung, P. H. J. & Batchelder, D. N. 1984 The study of model
polydiacetylene epoxy composites. 1. The axial strain in the fiber. J. Mater. Sci. 19, 3640–
3648.
Gommans, H. H., Alldredge, J. W., Tashiro, H., Park, J., Magnuson, J. & Rinzler, A. G. 2000
Fibers of aligned single-walled carbon nanotubes: polarized Raman spectroscopy. J. Appl.
Phys. 88, 2509–2514.
Haggenmueller, R., Gommans, H. H., Rinzler, A. G., Fischer, J. E. & Winey, K. I. 2000 Aligned
single-wall carbon nanotubes in composites by melt processing methods. Chem. Phys. Lett.
330, 219–225.
Harris, P. J. F. 1999 Carbon nanotubes and related structures. Cambridge University Press.
Hull, D. 1981 An introduction to composite materials. Cambridge University Press.
Iijima, S. 1991 Helical microtubules of graphitic carbon. Nature 354, 56–58.
Kearns, J. K. & Shambaugh, R. L. 2002 Polypropylene fibers reinforced with carbon nanotubes.
J. Appl. Polym. Sci. 86, 2079–2084.
Kelly, A. & Tyson, W. R. 1965 Tensile properties of fiber-reinforced metals—copper/tungsten
and copper/molybdenum. J. Mech. Phys. Solids 13, 329.
Kim, K. & Jin, J. 2001 Preparation of PPV nanotubes and nanorods and carbonized products
derived therefrom. Nano Lett. 1, 631–636.
Kroto, H. W., Heath, J. R., O’Brien, S. C., Curl, S. C. & Smalley, R. E. 1985 C-60: buckmin-
sterfullerene. Nature 318, 162–163.
Lillehei, P. T., Park, C., Rouse, J. H. & Siochi, E. J. 2002 Imaging carbon nanotubes in high
performance polymer composites via magnetic force microscopy. Nano Lett. 2, 827–829.
Lourie, O. & Wagner, H. D. 1998 Evaluation of Young’s modulus of carbon nanotubes by micro-
Raman spectroscopy. J. Mater. Res. 13, 2418–2422.
Matthews, M. J., Pimenta, M. A., Dresselhaus, G., Dresselhaus, M. S. & Endo, M. 1999 Origin
of dispersive effects of the Raman D band in carbon materials. Phys. Rev. B 59, R6585–6588.
Nesterenko, A. M., Kolesnik, N. F., Akhmatov, Y. S., Sukhomlin, V. I. & Pri-lutski, O. V. 1982
Metals 3. UDK 869.173.23. In News of the Academy of Science, USSR, pp. 12–16.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

Raman spectroscopy of CNT-based composites 2423

Nielsen, A. S. & Pyrz, R. 1999 Study of the influence of thermal history on the load transfer effi-
ciency and fibre failure in carbon/polypropylene microcomposites using Raman spectroscopy.
Compos. Interf. 6, 467–482.
Pócsik, I., Hundhausen, M., Koós, M. & Ley, L. 1998 Origin of the D peak in the Raman
spectrum of microcrystalline graphite. J. Non-cryst. Solids B 227–230, 1083.
Poulin, P., Vigolo, B. & Launois, P. 2002 Films and fibers of oriented single wall nanotubes.
Carbon 40, 1741–1749.
Qian, D., Dickey, E. C., Andrews, R. & Rantell, T. 2000 Load transfer and deformation mech-
anisms in carbon nanotube-polystyrene composites. Appl. Phys. Lett. 76, 2868–2870.
Rinzler, A. G. (and 14 others) 1998 Large-scale purification of single-wall carbon nanotubes:
process, product, and characterization. Appl. Phys. A 67, 29–37.
Robinson, I. M., Zakikhani, M., Day, R. J., Young, R. J. & Galiotis, C. 1987a Strain dependence
of the Raman frequencies for different types of carbon-fibers. J. Mater. Sci. Lett. 6, 1212–1214.
Robinson, I. M., Young, R. J., Galiotis, C. & Batchelder, D. N. 1987b Study of model polydi-
acetylene epoxy composites. 2. Effect of resin shrinkage. J. Mater. Sci. 22, 3642–3646.
Saito, R., Takeya, T., Kimura, T., Dresselhaus, G. & Dresselhaus, M. S. 1998 Raman intensity
of single-wall carbon nanotubes. Phys. Rev. B 57, 4145.
Salvetat, J. P., Bonard, J. M., Thomson, N. H., Kulik, A. J., Forro, L., Benoit, W. & Zuppiroli,
L. 1999 Mechanical properties of carbon nanotubes. Appl. Phys. A 69, 255–260.
Schadler, L. S., Giannaris, S. C. & Ajayan, P. M. 1998 Load transfer in carbon nanotubes epoxy
composites. Appl. Phys. Lett. 73, 3842–3844.
Shaffer, M. S. P. & Windle, A. H. 1999 Fabrication and characterization of carbon nano-
tube/poly(vinyl alcohol) composites. Adv. Mater. 11, 937–941.
Stephan, C., Nguyen, T. P., de la Chapelle, M. L., Lefrant, S., Journet, C. & Bernier, P. 2000
Characterization of single-walled carbon nanotubes-PMMA composites. Synth. Metals 108,
139–149.
Thomsen, C. & Reich, S. 2000 Double resonant Raman scattering in graphite. Phys. Rev. Lett.
85, 5214–5217.
Treacy, M. M. J., Ebbesen, T. W. & Gibson, J. M. 1996 Exceptionally high Young’s modulus
observed for individual carbon nanotubes. Nature 381, 678–680.
Valentini, L., Biagiotti, J., Kenny, J. M. & Santucci, S. 2003 Morphological characterization of
single-walled carbon nanotubes-PP composites. Compos. Sci. Tech. 63, 1149–1153.
Venkateswaran, V. D., Rao, A. M., Richter, E., Menon, M., Rinzler, A., Smalley, R. E. &
Eklund, P. C. 1999 Probing the single-wall carbon nanotube bundle: Raman scattering under
high pressure. Phys. Rev. B 59, 10 928–10 934.
Wagner, H. D., Amer, M. S. & Schadler, L. S. 2000 Residual compression stress profile in high-
modulus carbon fiber embedded in isotactic polypropylene by micro-Raman spectroscopy.
Appl. Compos. Mater. 7, 209–217.
Wood, J. R., Frogley, M. D., Meurs, E. R., Prins, A. D., Peijs, T., Dunstan, D. J. & Wagner, H. D.
1999 Mechanical response of carbon nanotubes under molecular and macroscopic pressures.
J. Phys. Chem. B 103, 10 388–10 392.
Wood, J. R., Zhao, Q., Frogley, M. D., Meurs, E. R., Prins, A. D., Peijs, T., Dunstan, D. J. &
Wagner, H. D. 2000 Carbon nanotubes: from molecular to macroscopic sensors. Phys. Rev.
B 62, 7571–7575.
Wood, J. R., Zhao, Q. & Wagner, H. D. 2001 Orientation of carbon nanotubes in polymers and
its detection by Raman spectroscopy. Composites A 32, 391–399.
Yakobson, B. I. & Smalley, R. E. 1997 Fullerene nanotubes: C1 000 000 and beyond. Am. Sci. 85,
324–337.
Young, R. J., Thongpin, C., Standford, J. L. & Lovell, P. A. 2001 Fragmentation analysis of
glass fibres in model composites through the use of Raman spectroscopy. Composites A 32,
253–269.

Phil. Trans. R. Soc. Lond. A (2004)


Downloaded from rsta.royalsocietypublishing.org on May 23, 2011

2424 Q. Zhao and H. D. Wagner

Zhao, Q. & Wagner, H. D. 2003 Two-dimensional stress field mapping of fiber reinforced polymer
composites using carbon nanotube strain sensors. Composites A 34, 1219–1225.
Zhao, Q., Frogley, M. D. & Wagner, H. D. 2001a The use of carbon nanotubes to sense matrix
stresses around a single glass fiber. Compos. Sci. Tech. 61, 2139–2143.
Zhao, Q., Wood, J. R. & Wagner, H. D. 2001b Using carbon nanotubes to detect polymer
transitions. J. Polym. Sci. B 39, 1492–1495.
Zhao, Q., Wood, J. R. & Wagner, H. D. 2001c Stress fields around defects and fibers in a polymer
using carbon nanotubes as sensors. Appl. Phys. Lett. 78, 1748–1750.

Phil. Trans. R. Soc. Lond. A (2004)

View publication stats

You might also like