Saito 2011
Saito 2011
To cite this article: R. Saito , M. Hofmann , G. Dresselhaus , A. Jorio & M. S. Dresselhaus (2011)
Raman spectroscopy of graphene and carbon nanotubes, Advances in Physics, 60:3, 413-550,
DOI: 10.1080/00018732.2011.582251
REVIEW ARTICLE
This paper reviews progress that has been made in the use of Raman spectroscopy to study
graphene and carbon nanotubes. These are two nanostructured forms of sp2 carbon materi-
als that are of major current interest. These nanostructured materials have attracted particular
attention because of their simplicity, small physical size and the exciting new science they have
introduced. This review focuses on each of these materials systems individually and compara-
tively as prototype examples of nanostructured materials. In particular, this paper discusses the
power of Raman spectroscopy as a probe and a characterization tool for sp2 carbon materials,
with particular emphasis given to the field of photophysics. Some coverage is also given to the
close relatives of these sp2 carbon materials, namely graphite, a three-dimensional (3D) mate-
rial based on the AB stacking of individual graphene layers, and carbon nanoribbons, which
are one-dimensional (1D) planar structures, where the width of the ribbon is on the nanometer
length scale. Carbon nanoribbons differ from carbon nanotubes is that nanoribbons have edges,
whereas nanotubes have terminations only at their two ends.
Contents PAGE
This review of the Raman spectroscopy of graphene and carbon nanotubes starts with a brief
presentation of background material to set the stage for a discussion of the present status of
knowledge on this topic by providing a broad overview of the field in Section 1. This is followed,
in Section 2, by an overview discussion of experimental progress that has been made in recent years
using many experimental techniques. Advances in the theoretical understanding needed to interpret
the many new results that appear daily on the photo-physics of graphene and carbon nanotubes
are reviewed in Section 3. The detailed consideration of the Raman spectroscopy of graphene
and carbon nanotubes is presented in Sections 4 and 5, both individually and comparatively,
while Section 6 looks at future developments in this field. A textbook helpful for understanding
background material relevant to this review article has recently been published [1].
Hereafter in, Section 1, we provide background to Raman spectroscopy in general, giving
special attention to resonance Raman spectroscopy, with illustrations of the various photo-physical
phenomena, which are given in terms of graphene and carbon nanotubes. Raman measurements in
low-dimensional (one- and two-dimensional) systems are then discussed in general terms, giving
specific examples of the differences and similarities between the Raman spectra for graphene and
carbon nanotubes in relation to other sp2 carbons. Finally, we discuss what can be learned from the
Raman spectrum from one laser line in comparison to what can be learned from using a continuum
of laser lines.
1. Basic concepts
In this section, we introduce the basic concepts of Raman spectroscopy, starting with where they
stand within the broad picture of light–matter interactions and then considering their general char-
acteristics. The concepts of first-order and higher-order scattering processes, Stokes vs. anti-Stokes
processes, lineshapes, resonance and coherent processes are introduced, as well as the classical
treatment of the Raman effect. Subsequently, we present a historical description of the Raman
spectroscopy used to study and characterize graphitic materials, from graphite to single-walled
carbon nanotubes (SWNTs) and graphene. This brief introduction to the Raman spectroscopy of
sp2 nanocarbons should be useful to general readers who may not be so familiar with Raman
spectroscopy.
1.1.1. Graphite
Carbon materials have been the objects of study and use for many years [2]. Three-dimensional
(3D) graphite is one of the longest known forms of pure carbon, naturally occurring on the surface of
the earth as a mineral (as, for example, in the mineral deposits of Ticonderoga (NewYork) graphite,
Ceylon (Sri Lanka) and the large graphite deposits in Minas Gerais, Brazil). Structurally, graphite
is a layered material, with individual graphene layers, as shown in Figures 1 and 2(a). These
individual graphene layers are stacked in the ABAB Bernal stacking order in the most common
form of graphite, as shown in Figure 2(b) to illustrate the relative in-plane atomic arrangements
of the A and B carbon atoms within the layer plane and in adjacent layers. Here, one type of
carbon atom (labeled A) is aligned on top of another A atom in the direction perpendicular to the
graphene layer, while the other type (the B carbon atom) is aligned in every other layer1 [2], so
that the graphene planes are arranged in the so-called ABAB Bernal stacking sequence, as shown
in Figure 2(c) [6], where a carbon atom is found on adjacent layer planes over the empty center of
a hexagon. This AB stacking order (Figure 2(c)) also applies to bi-layer graphene prepared by the
418 R. Saito et al.
Figure 1. STM image of graphite. Notice the different brightness for the A and B atoms (see footnote 1).
Reprinted from Carbon, 48(5), M.M. Lucchese et al. pp. 1592–1597 [3]. Copyright © (2010) Elsevier.
mechanical exfoliation method [7,8], while Figure 2(d) shows the stacking arrangement of trilayer
graphene. In graphite, layers 1 and 3 are crystallographically equivalent and are translated from
one another by the out-of-plane lattice parameter c = 0.670 nm (see Figure 2(d)), thus generating
hexagonal graphite. Another related crystalline arrangement is the less common form known as
rhombohedral graphite [9], which has an ABC stacking order, consisting of three layers and a
lattice constant of 1.005 nm.
Of all materials, graphite has the highest melting point (4200 K), the highest thermal conduc-
tivity (3000 W/mK) and a high room-temperature electron mobility (30, 000 cm2 /V s) [10,11]. 3D
graphite was synthesized for the first time in 1960 by Arthur Moore and co-workers [12–16] and
their high-temperature, high-pressure synthesis method yielded the material commonly known as
highly oriented pyrolytic graphite (HOPG). Graphite and its related carbon fibers [17–19] have
been used commercially for decades [20]. Carbon fiber applications range from use as conductive
fillers, and as mechanical structural reinforcements in composites (e.g., in the aerospace industry),
to their use as electrode materials for making steel, exploiting their good electrical conductivity
and in lithium ion battery applications exploiting their high resiliency [20,21].
1.1.2. Fullerenes
In 1985, the discovery of another unique sp2 carbon system took place, the observation of the
C60 fullerene molecule [22]. The fullerene molecule consists of 60 carbon atoms with mostly sp2
bonding and appropriate π bonding to form a closed surface with full icosahedral symmetry.
Because of topological restrictions, fullerenes, in general, have 12 pentagonal rings and any
numbers of hexagonal rings, thereby generating a large variety of Cn fullerene molecules. The C60
molecule with full icosahedral symmetry can be regarded as the first isolated carbon nanosystem.
Fullerenes stimulated and motivated a large scientific community into new research directions from
the time of their discovery up to the end of the twentieth century, but fullerene-based applications
Advances in Physics 419
Figure 2. (a) Top view of the real space unit cell of mono-layer graphene showing the inequivalent carbon
atoms A and B and the graphene unit vectors a1 and a2 . (b) Top view of the real space bi-layer graphene
structure. The light/dark gray dots and black circles/black dots represent the carbon atoms in the upper and
lower layers, respectively, of bi-layer graphene (2-LG). (c) The unit cell and the x and y unit vectors of bi-layer
graphene and (d) the same as (c) but for trilayer graphene. (e) The reciprocal space unit cell showing the first
Brillouin zone with its high symmetry points and lines, such as the line T connecting to K. (f) The Brillouin
zone for 3D graphite showing high symmetry points and axes. Here is a high symmetry point along the
axis connecting points A and , and u is a general point in the KM plane. Adapted figure with permission
from A. Jorio et al. Spectroscopy in Graphene Related System, 2010 [1] Copyright © Wiley-VCH Verlag
GmbH & Co. KGaA; and with permission from L. M. Malard et al., Physical Review B 79, p. 125426, 2009
[4]. Copyright © (2009) by the American Physical Society; and with permission from Physics Reports 473,
L.M. Malard et al. pp. 51–87 [5]. Copyright © (2009) Elsevier.
remain sparse to date. In this review, we do not mention fullerenes much. See [23] for more
information on fullerenes.
Figure 3. Early transmission electron microscopy images of carbon nanotubes [29]. The early reported
observations of nanotubes (a) in 1952 [28] and (b) in 1976 [25]. (c) Observation of SWNTs that launched
the field in 1993 [30,31] together with an example of their observation. Adapted figure with permission
from Carbon 14(2), A. Oberlin et al., pp. 133–135 [25]. Copyright © (1976) Elsevier; and with permission
from Carbon 44, M. Monthioux and V.L. Kuznetsov, pp. 1621–1623 [29]. Copyright © (2006) Elsevier; and
with permission from Macmillan Publishes Ltd. Nature [30]. Copyright © (1993); and with permission from
Macmillan Publishes Ltd. Nature [31], Copyright © (1993).
amount of millimeter-long nanotubes with nearly 100% SWNT purity (absence of other carbon
forms) have now been achieved [34], and further improvements in nanotube synthesis are evolving
rapidly at this time. Substantial success with the separation of nanotubes by their (n, m) structural
indices, metallicity (semiconducting or metallic) and by their length has been achieved by different
methods, especially by the density gradient approach of Hersam and Arnold [35]. Advances have
been made with doping either n-type or p-type nanotubes for the modification of their properties,
as summarized in reference [36,37]. Studies on nanotube mechanical properties [37,38], optical
properties [39–45], magnetic field-dependent properties [46], optoelectronics [47,48], transport
properties [49] and electrochemistry [50,51] have exploded, revealing many rich and complex
fundamental excitonic and other collective phenomena [21]. Nanotube-based quantum transport
phenomena, including quantum information applications, spintronics and superconducting effects,
have also been explored [49]. After more than a decade and a half of intense activity in carbon
nanotube research, more and more attention is now being focussed on the practical applications of
the many unique and special properties of carbon nanotubes [20]. Further background information
on the synthesis, structure, properties and applications of carbon nanotubes can be found in [1,21]
and some of these topics are further emphasized in the present review.
1.1.4. Graphene
The interest in a single atomic layer of sp2 carbon (called graphene; see Figure 1) goes back to
the pioneering theoretical work of Wallace in 1947 [52] which was for many years used as a
model system for all sp2 carbons. This very early work provides a framework for comparing the
structures of graphite, fullerenes, carbon nanotubes and other sp2 nanocarbons. The synthesis of
single-layer graphene was actually reported by Boehm in 1962 [53], but this early discovery was
neither confirmed nor followed up for many years.
More recently, mono-layer graphene was synthesized from nano-diamonds in 2001 [54] and
from SiC [55]. The material synthesized from nano-diamonds is generally a few-layer graphene
material but thin ribbon specimens of mono-layer thickness are also contained in such samples. In
the studies on these nanoribbons, emphasis was given to the properties of edges of the nanoribbons
and especially to the magnetic properties of these edges [54,56–59]. The graphene prepared from
heating SiC to 1300 ◦ C emphasized the 2D electron gas properties of this graphene in an electric
field, but did not especially focus on the number of layers [55].
The widespread study of graphene was launched by the preparation of mono-layer graphene
by Novoselov et al. [7], using a simple Scotch tape method to prepare and transfer mono-layer
graphene from the c-face of graphite to a suitable substrate such as SiO2 for the measurement
of the electrical and optical properties of mono-layer graphene [60]. The Novoselov and Geim
studies of transport in few-layer graphene in 2004 [7] led to a renewed interest in mono-layer and
Advances in Physics 421
few-layer graphene and to an in-depth study of the unique properties of this material in the mono-
layer and bi-layer limit. Surprisingly, this very basic system, which had been conceptually utilized
by researchers over a period of many decades, suddenly appeared on the experimental scene,
demonstrating many novel physical properties that were not even imagined previously [60,61].
The discovery of these novel properties launched a rush into the study of graphene science in the
first decade of the twenty-first century, and culminating in the 2010 Nobel Prize in physics [62].
Besides the outstanding mechanical properties [63] (breaking strength ∼40 N/m, Young’s
modulus ∼1.0 TPa) and thermal properties [64,65] (room temperature thermal conductivity
∼3000 W m−1 K −1 [65]), the scientific interest in graphene was stimulated [66,67] by the
widespread report of the relativistic (massless) electronic properties of the conduction elec-
trons (and holes) in a single layer less than 1 nm thick, with a state-of-the-art mobility reaching
μ = 200, 000 cm2 /V s at room temperature for freely suspended graphene [66–71]. Other unusual
properties have been predicted and demonstrated experimentally, such as the minimum conductiv-
ity and the half-integer quantum Hall effect in mono-layer graphene [72] and the integer quantum
Hall effect in bi-layer graphene [73], ambipolar transport by either electrons or holes by the vari-
ation of a gate potential, operation as a transparent conductor [47,74], Klein tunneling [75–82],
negative refractive index and Veselago lensing [80], anomalous Andreev reflections at metal–
superconductor junctions [76,81–84], anisotropies under antidot lattices [85] or periodic potentials
[86], and a metal–insulator transition via hydrogenation of graphene [87]. Applications such as
fillers for composite materials, as super-capacitors, batteries, interconnects and field emitters are
being developed, although it is still too early to say to what extent graphene will be able to com-
pete with carbon nanotubes and other established materials systems in the applications world
[88]. Although nanotubes and graphene are both carbon-based nanostructures, they have different
properties related to the planar aspects of graphene and the tubular aspects of nanotubes, and this
basic difference should distinguish their optimal usage in applications.
1.1.5. Nanoribbons
Graphene nanoribbons are of particular interest for introducing a bandgap into graphene-related
systems. Bandgaps are needed for many electronics applications of nanomaterials. Since graphene
can be patterned using, for example, high-resolution lithography [55,89], nanocircuits with
graphene–nanoribbon interconnects can be fabricated. Many groups are now fabricating devices
using graphene and also graphene nanoribbons, which have a long length and a small nanoscale
width, and where the ribbon edges play an important role in both determining their electronic
structure and exhibiting unusual spin polarization properties [56]. Nanoribbons of small widths
exhibit 1D behavior analogous to carbon nanotubes, but have a high density of electronic states
at the Fermi level for the case of well-defined zigzag edges. This high density of electronic states
allows us to experimentally distinguish zigzag nanoribbons from armchair and chiral nanoribbons
which do not exhibit this property. While lithographic techniques have limited resolution for the
fabrication of small width nanoribbons (<20 nm wide), chemical [90] and synthetic [91] methods
have been employed successfully, including the unzipping of SWNTs as a route to produce carbon
nanoribbons [92,93]. Carbon nanoribbons have been shown to be especially sensitive for the study
of edge structures and edge properties and edge reconstruction effects [94].
Figure 4. Raman spectra from several sp2 nanocarbon and bulk carbon materials. From top to bottom:
crystalline mono-layer graphene, HOPG, an SWNT bundle sample, damaged graphene, single-wall carbon
nanohorns (SWNH). The most intense Raman peaks are labeled in a few of the spectra [1,96]. Note that
some authors call the G’ by 2D and the G” by 2D’ [97]. Reprinted with permission from M.S. Dresselhaus
et al., Nano Letters 10, pp. 953–973, 2010 [96]. Copyright © (2010) American Chemical Society.
we can obtain the phonon frequency which is useful for identifying the origin of an unknown
structure of a newly discovered molecule or of a new material in chemistry [98]. Among all
possible phonon modes for sp2 carbons, only a limited number of phonons are Raman-active
modes (namely those with A, E1 and E2 symmetry for carbon nanotubes, and E2g for graphite)
[99–101]. A common metric used to characterize the defect density in a material is the ratio of the
intensities of the disorder-induced D-band to the symmetry-allowed G-band ratio (ID /IG ) [102].
Study of the D- and G-band modes by Raman spectroscopy (see Figure 4) yields information
about the crystal structure of the material and about many of its interesting physical properties.
Raman spectroscopy for the various sp2 carbon materials (see Figure 4) has been mainly
used for sample characterization and these different carbon materials exhibit characteristic dif-
ferences related to the small differences in their structures. The fundamental sp2 carbon material
is mono-layer graphene which has the simplest and most fundamental spectrum showing the two
Raman-allowed features that appear in all sp2 carbon materials – the first-order G-band and the
second-order symmetry-allowed G -band,2 where the symbol G is used to denote “graphitic.” The
next most commonly observed feature is the D-band that is a defect-activated Raman mode. The
D-band occurs at about 1350 cm−1 at 2.41 eV laser excitation energy (Elaser ) and is highly disper-
sive as a function of Elaser (see Section 2.8.9). Since the graphite melting temperature is very high
(over 4200 K) and since no actual carbon materials are defect-free, the D/G-band intensity ratio
(ID /IG ) provides a sensitive metric for the degree of disorder in sp2 carbon materials over a wide
temperature range. In the case of fullerenes, a special Raman-active phonon mode related to the
vibrations of a pentagonal ring (1469 cm−1 ) is particularly sensitive for understanding the molec-
ular structure [23]. In the case of carbon nanotubes, it is common to survey unknown samples
using Raman spectroscopy to check for the presence of nanotubes in the sample by observing the
cylindrical-specific, Raman-active-mode radial breathing mode (RBM) in which atoms around the
circumference of a single wall carbon nanotube (SWNT) are vibrating in a breathing mode in the
radial direction (see Figure 4) [32,100,103–110]. This vibrational mode is unique to carbon nan-
otubes and serves to sensitively identify their presence in a given sample. Since the RBM frequency
Advances in Physics 423
ωRBM is inversely proportional to the nanotube diameter dt , we can thus estimate the diameter
distribution of the nanotubes that are contained in a given sample [111]. When we observe an
isolated nanotube, we can use its Raman spectrum to obtain its detailed structure, which involves
identification of the spatial orientation, the diameter and the chiral angle of the nanotube, as well
as the nanotube (n, m) chirality assignment. This (n, m) assignment is based on the concept of
the resonance Raman effect (see Section 1.5.2). In the case of graphene [112–114], the inten-
sity ratio (IG /IG ) and the lineshape of the G -band (along with other indicators) can be used for
identifying the number of graphene layers (see Section 4.2.1). For graphene ribbons, we can use
Raman spectroscopy to study the edge structure of the ribbons [94,115] to yield information about
the structure and properties of graphene ribbons (see Section 4.4). For all these reasons, Raman
spectroscopy is very sensitive for the characterization of sp2 carbons.
Since most carbon materials are not soluble in water, Raman spectroscopy is useful as an
in situ, non-contact, non-destructive measurement tool that can be used at room temperature
and under ambient operating conditions, as well as for freely suspended carbon nanotubes and
graphene samples. Combining the continuing advances in optical techniques with new theoretical
developments that are rapidly developing, Raman spectroscopy studies of graphene and carbon
nanotubes have provided a great deal of information about their solid-state properties, which are
the main subjects of this review, including their behavior as a function of temperature, pressure
and Fermi energy [1].
Figure 5. Photo-physical mechanisms operative for various regions of the electromagnetic spectrum as
photons in various energy ranges interact with materials. Reprinted figure with permission from A. Jorio et
al. Spectroscopy in Graphene Related Systems, 2010 [1]. Copyright © Wiley-VCH Verlag GmbH & Co.
KGaA.
have excitonic levels of a few meV below the bandgap, the excitonic levels in carbon nanotubes are
much larger (on the order of a few hundred meV), emphasizing the greater importance of excitonic
effects in low-dimensional nanosystems. Optical absorption in the case of carbon nanotubes was
predicted by Ando to be excitonic as far back as 1997 [120] and the importance of excitons
in nanotubes was demonstrated experimentally by the two photon experiments carried out by
the Heinz group in 2005 [121] and a similar result was independently obtained by the Berlin
group [122]. These experiments confirm theoretical concepts that excitonic effects would be
enhanced in low-dimensional systems. Although excitonic effects were found to be much larger
in semiconducting nanotubes than in metallic nanotubes, excitons nevertheless have been found
in these works to also dominate optical absorption processes in metallic nanotubes.
Figure 6. The light–matter interaction, showing the most commonly occurring processes. The waved arrows
indicate incident and scattered photons. The vertical arrows denote photon-induced transitions between
(a) vibrational levels and (b–e) electronic states. Curved arrow segments indicate electron–phonon (el–ph)
(hole–phonon) scattering events. In (e), the shortest vertical arrow also indicates an el–ph transition in Raman
scattering. In (d, e), the processes are resonant if the incident (or scattered) photon energy exactly matches the
energy difference between initial and excited electronic states. When far from the resonance window where
resonance occurs, the optical transition is called a virtual transition. The intensity for resonance Raman
scattering can be much larger for vertical processes that are resonant than those that are not resonant [1,95].
Reprinted figure with permission from A. Jorio et al. Spectroscopy in Graphene Related Systems, 2010 [1].
Copyright © Wiley-VCH Verlag GmbH & Co. KGaA.
(el-ph) interaction that occurs in the Raman scattering process. In this process the electron is
excited to a virtual state, for which the stable geometry of the chemical bond is no longer iden-
tical to that of the ground state. This perturbation generates a force resulting in atomic motion,
thereby providing a quantum explanation for the el–ph interaction. If the incident photons are
introduced through a sufficiently short light pulse with a duration comparable to the frequency
of the phonon, then all atoms start to move at the same time as a result of the el–ph interaction,
thereby creating a coherent motion of the atoms which can be detected by a second pulse of light
at a frequency at which the material is transparent. This sensitive technique is known as coherent
phonon spectroscopy.
bottom (top) of the conduction (valence) energy band by creating multiple phonons of different
frequencies through el–ph coupling, in which the phonons are selected such that the initial and
final electronic states satisfy both energy and momentum conservation requirements. In the case
of a metal, such photo-excited electrons (together with their holes) will decay down (up) to
the ground states without emitting a photon, and such processes are called non-radiative decay
process as shown in Figure 6(b). In graphene or metallic carbon nanotubes, non-radiative decay
that generates heat frequently occurs and can have a special character because of their linear E(k)
dispersion relation.
an Indian scientist who was awarded the Nobel Prize in Physics in 1930 for his work on “the
scattering of light and for the discovery of the effect named after him”. In the Raman process,
an incident photon with energy Ei = Elaser and momentum ki = klaser reaches the sample and is
scattered, resulting in a photon with a different energy Es and a different momentum ks . For energy
and momentum conservation,
Es = Ei ± Eq and ks = ki ± q, (1)
where Eq and q are the energy and momentum change during the scattering event induced by
electromagnetic excitation of the medium. The quantities Eq and q can be considered to be the
energy and the momentum of the phonon.
The Raman process which emits (absorbs) a phonon is called a first-order Raman process. In
order to recombine an electron at ks with a hole at ki , the wavevector q should be almost zero.
Thus, only phonons near the point (zone-centered phonon) in the phonon dispersion relation can
be Raman-active modes. However, when we consider second-order Raman processes in which
two scattering events are involved, the restriction for q = 0 is relaxed. Further, the photo-excited
electrons in sp2 carbons are located in k space near the hexagonal corners of the 2D Brillouin zone
(BZ), named the K and K points, where the states of lowest energy are located. Here, there are
two possibilities for q = 0 scattering: intra-valley (K → K, K → K ) and inter-valley scattering
(K → K , K → K). We will show (see Section 2.8.4) that a double resonance Raman process
involving excitations near the K and K points in the 2D BZ yields a large Raman signal.
in Section 1.4.7, the characteristics of the Raman excitation profile, particularly the resonance
window width.
Figure 7. Schematics showing the Rayleigh line (at 0 cm−1 ) and the Raman spectrum. The Rayleigh intensity
is always much stronger and it has to be filtered out for any meaningful Raman experiment. The Stokes process
(positive frequency peaks) are usually stronger than the anti-Stokes process (negative frequency peaks) due to
phonon creation/annihilation statistics. Reprinted from Carbon, 48(5), M.M. Lucchese et al. pp. 1592–1597
[3]. Copyright © (2010) Elsevier.
the damping or the energy uncertainty or the phonon lifetime. The damping of the amplitude (as
characterized by q ) is observed as Elaser is tuned (and thus as the scattered light energy is varied).
The damping of I(ω) thus provides information on the phonon lifetime, t. The uncertainty
principle Et ∼ gives an uncertainty in the value of the phonon energy, as measured in the
Raman spectrum, which corresponds to the spectral FWHM of 2q . Therefore, q is the inverse of
the lifetime for a phonon, and Raman spectra in this way provide information on phonon lifetimes.
There are two origins for the finite phonon lifetime: the anharmonic potential and the el–ph
interaction, each of which are discussed below. This is followed by a further discussion of the
lineshapes I(ω) observed for phonons.
(i) Anharmonic potential
Anharmonicity of the inter-atomic potential for the phonon occurs for large r far from the
potential minimum. In this regime, the wave vector q of the phonon is no longer a good quantum
number and phonon scattering occurs by emitting a phonon (third-order process) or by phonon–
phonon scattering (fourth-order anharmonicity). Anharmonicity gives the main contribution to
the thermal expansion process (third-order process) and to the thermal conductivity (fourth-order
process).
(ii) Electron–phonon interaction
Another possible interaction is the el–ph interaction in which a phonon excites an electron in
the valence band to the conduction band or scatters a photo-excited electron to other unoccupied
states. The former el–ph process works for electrons in the valence band, while the latter el–ph
process works for electrons in excited states. Thus, the origins of the finite lifetime of the phonon
are different from each other for the case of electron and hole excitation, and is one mechanism
for breaking the symmetry between electrons and holes in graphene.
contribution. Then the Raman peak will be a convolution of several Lorentzian peaks, depending
on the frequency and weight of each phonon contribution.
One case of importance occurs when the lattice vibration couples to free electrons, as occurs
in graphene or metallic nanotubes when an el–ph interaction takes place. In this case, additional
line broadening and even distorted (asymmetric) lineshapes can result, and this effect is known as
the Kohn anomaly (KA) [127]. In cases where phonons are coupled to the continuum excitation
spectra of free electrons, the Raman peak may exhibit a so-called Breit–Wigner–Fano (BWF)
lineshape, given by [128,129]
where 1/qBWF is a measure of the interaction of a discrete level (the phonon) with a continuum
of states (the electrons). Here ωBWF is the BWF peak frequency at the maximum intensity I0 , and
BWF is the frequency half width-half maximum for the intensity profile of the BWF peak. Such
effects are observed in certain metallic sp2 carbon materials and are discussed later, in connection
with metallic carbon nanotubes.
1
I0 (ω) ∝ , (6)
[ω − ω(q0 )]2 + [0 /2]2
as described above. A disordered distribution of point defects as would be produced by ion implan-
tation, however, will scatter phonons and will also add a contribution to the FWHM by coupling
phonons of wavevector q0 to those of wavevector q0 + δq [130]. In the limit of low levels of
disorder, the coupling will be most effective for small δq, so the phonon wave packet in k-space
can be described by a Gaussian function exp[−(q − q0 )2 Lpc 2
/4] centered at q0 and having a width
proportional to 1/Lpc ≈ δq. Therefore, in real space Lpc is a measure of the phonon coherence
length, which should also be a good measure of the average distance between point defects. Then,
the Raman intensity for the disordered graphene I(ω) can be written as [130–134]
W (q) exp −(q − q0 )2 Lpc
2
/4
I(ω) ∝ d2q , (7)
BZ [ω − ω(q)]2 + [0 /2]2
where the integral is taken over the 2D BZ of graphene and W (q) is a weighting function that
describes the wavevector dependence of the el–ph coupling for the Raman process.
where the resonant lineshape I(Elaser ) consists of two peaks at Elaser = Eii (incident resonance
condition) and Elaser = Eii ± Eq (the scattered resonance condition). Here γr is the FWHM width
discussed below. Experimentally it is not always possible to resolve the observed I(Elaser ) lineshape
into two peaks.
The FWHM width of each peak in the Raman excitation profile is the resonance window width
γr , and is related to the lifetime of an electron in its excited states. If the lifetime of the photo-
excited electron is finite, then the resonance condition in the Raman excitation profile may show
departures from Equation (8) which assumes γr Eq . The finite lifetime of the photo-excited
electron is subject to the uncertainty relation Et ∼ .
The photo-excited carrier can be relaxed from the excited states by several processes, each
occurring according to different time scales:
When we consider the Coulomb interaction for a given electron, the other electrons should excite
the first electron to an unoccupied state. Thus, the Coulomb interaction depends on the metallicity
of the material. In the case of carbon nanotubes, the interaction between two excitons that is
relevant to this term is known as the Auger process. On the other hand, a photo-excited electron
has a definite lifetime for emitting any energy–momentum conserved phonon. Thus, the el–ph
interaction of the electron from a state k (a photo-excited state) to the energy–momentum conserved
k + q (phonon emitting electron state) is important. Note that γr (the resonance window width,
Eq. (8)) is physically different from q (the Raman spectral width, Eq. (4)).
Figure 8. Typical electronic density of states for 3D, 2D, 1D and 0D systems.
434 R. Saito et al.
1.5.1. Cutting lines and van Hove singularities of the density of states
When the 2D sheets of graphene are rolled up to form 1D nanotubes, different subbands in the 1D
reciprocal space of the nanotube can be extended into the 2D reciprocal space of a single sheet of
the parent bulk layered material as a set of parallel equi-distant cutting lines [32,135,136]. This
procedure is shown in Figure 9(a) for states near the K point.
Figure 9(b) shows the electronic density of states (DOS) related to the nanotube electronic
band structure plotted schematically in Figure 9(a). Each of N cutting lines in Figure 9(a) (except
for the one that crosses the degenerate K point) gives rise to a local maximum in the DOS g(E) in
Figure 9(b), known as a (1D) van Hove singularity (vHS), given by
N
2 ∂Eμ (k) −1
g(E) = δ[Eμ (k) − E] dk. (9)
N μ=1 ∂k
The four vHSs in Figure 9(b) are labeled by Ei(v) and Ei(c) for the electronic subbands in the valence
and conduction bands, correspondingly. The presence of vHSs in the DOS of 1D structures makes
these structures behave differently from their related 3D and 2D counterpart materials, as can be
seen in Figure 8. A finite density of states between the first singularities in the valence band and
conduction band for metallic nanotubes is shown in Figure 9(b).
Figure 9. (a) The energy–momentum contours for the valence and conduction bands for a 2D system, with
each band obeying a linear dependence for E(k) and forming a degenerate point K where the valence and
conduction bands touch to define a zero gap semiconductor. The cutting lines of these contours denote the
dispersion relations for the 1D system derived from the 2D system. Each cutting line gives rise to a different
energy subband. The energy extremum Ei for each cutting line at the wave vector ki occurs at a van Hove
(v) (c)
singularity. The energies Ei and Ei for the valence and conduction bands and the corresponding wave
(v) (c)
vectors ki and ki at the van Hove singularities are indicated on the figure by the solid dots. (b) The 1D
density of states (DOS) for the conduction and valence bands in (b) corresponding to the E(k) dispersion
relations for the 1D subbands shown in (a) as thick curves. The DOS shown in (b) is for a metallic 1D system,
because one of the cutting lines in (a) crosses the degenerate Dirac point (the K point in the graphite Brillouin
Zone (BZ)). For a semiconducting 1D system, no cutting line crosses the degenerate point, thus resulting in
(v) (c)
a band gap opening up in the DOS between the van Hove singularities E1 and E1 [135].
Advances in Physics 435
energy band extremum, g(E) is given by g ∝ (E − E0 )[(D/2)−1] , where D is an integer, denoting the
spatial dimension and D assumes the values 1, 2, and 3, respectively, for 1D, 2D, and 3D systems
[137]. The parameter E0 appearing in the density of states g(E) denotes the energy band minimum
(or maximum) for the conduction (valence) energy bands. For a 1D system, E0 would correspond
to the energy of a vHS in the DOS occurring at each subband edge, where the magnitude of the
DOS becomes very large. One can see from Figure 8 that 1D systems exhibit DOS profiles which
have some similarity to the case of 0D systems, with both 0D and 1D systems having very sharp
maxima at certain energies, in contrast to the DOS profiles for 2D and 3D systems, which show
a more monotonic increase with energy (see Figure 8). However, the 1D DOS is different from
the 0D DOS (δ function at each discrete energy level) in that the 1D DOS has a sharp threshold
and a decaying tail for each cutting line, so that the 1D DOS does not go to zero between the
sharp maxima, as the 0D DOS does (see Figure 8). This is even true for semiconducting nanotubes
which have a finite band gap and no occupied states between the first cutting lines in the valence
and conduction bands. The extremely high values of the DOS at the vHSs allow us to observe
physical phenomena for individual 1D nanostructures in various experiments, as discussed in
Section 1.1.6.
Figure 10. Qualitative comparative evaluation of the amount of information vs. the simplicity of performing
an optical experiment in SWNTs based on the experience of the authors. The position of each technique in the
plot is defined both by phyiscal limitations (e.g. photoluminescence is not available from metallic SWNTs)
and by aspects of practical implementation.
In Section 2, we review a number of optical and spectroscopic techniques that are used for
characterizing materials and especially focusing on sp2 carbon materials, carbon nanotubes and
graphene. We start this section with a brief introduction to the special properties of electrons and
phonons in graphene.
and increase as we move away from the K and K points. The anomalous and symmetric behavior
of the electrons and holes in graphene mainly originates from this unusual linear E(k) energy
dispersion relation of graphene near the Fermi energy, EF .
Electronic transitions occur from the electron-occupied π band to the electron-unoccupied
π ∗ energy band. Selection rules forbid intra-atomic transitions from 2p to 2p states. However,
inter-atomic transitions from 2p to 2p states are allowed both for the nearest-neighbor pairs of
carbon atoms and for further neighbors as well. Since the energy dispersion is linear near the K
and K points, the electronic dispersion in graphene forms Dirac cones. Optical transitions for a
given laser excitation energy Elaser occur on the equi-energy lines around the K and K points.
Because of the three-fold symmetry of E(k) around the K or K points, the equi-energy lines are
distorted into a triangle with increasing energy [32,138]. This trigonal warping effect of the energy
dispersion and of the Fermi surface dominate the electronic properties.
Figure 11. Cutting lines near the K point in the 2D BZ of graphene for (a) Type 0 (or Mod 0) metallic
SWNTs, (b) Type I (Mod 2) and (c) Type II (Mod 1) semiconducting SWNTs.
semiconducting nanotubes (S-SWNTs), respectively.9 In Figure 11, we show the geometry of the
cutting lines near the K point in the 2D BZ of graphene. In the case of type I (II) semiconducting
nanotubes, the K point lies at the one-third (two-thirds) position between two cutting lines, since
the distance of the K point from the cutting line at the point is given by (2n + m)/3 times K1 ,
where K1 is a reciprocal lattice vector in the direction of the chiral vector [32]. In Figure 11(a),
we see that mod(2n + m, 3) = 0 corresponding to metallic nanotubes. Each cutting line results in
an optical transition for its Eii value. Further discussion of this classification is given in [107,135,
136,141,145,146]. Because of the anisotropy of the effective mass tensor around the K point, the
optical properties of S-SWNTs depend on their Type I or II classification (see Figure 11(b) and
(c)). We call this classification the semiconductor type-dependence.1
Figure 12. (a) The optical absorption spectra from SWNTs and from colloidal graphite shown comparatively.
(b) Illustration of interband processes giving rise to the optical absorption peaks in semiconducting and
metallic SWNTs [42].
Advances in Physics 441
simple predictions for Dirac Fermions in idealized graphene as we move sufficiently far from the
Dirac point in k space.
Figure 13. (a) Schematic diagram for a Rayleigh scattering measurement. Microscope objectives focus the
incident light on a suspended nanotube and collect the radiation of the scattered light. Using a super-continuum
source, different wavelengths can be detected simultaneously using a spectrometer and a multichannel (CCD)
camera. (b) Electron micrograph of an individual suspended SWNT in a geometry used for Rayleigh scattering
spectroscopy. Adapted from Carbon Nanotubes: Advanced Topics in the Synthesis, Structure, Properties and
Applications, T.F. Heinz [44]. Copyright (2008) from Springer.
442 R. Saito et al.
Figure 14. The Rayleigh scattering spectra of two different (n, m) semiconducting SWNTs shown compar-
atively for type I(a), type II(b) semiconducting SWNTs. The peaks in both cases correspond to the ES33 and
ES44 interband transitions. The comparison in (a) corresponds to two nanotubes of the same (2n + m)mod3
type, but having different diameters; the comparison in (b) corresponds to two nanotubes of different type,
but with similar SWNT diameters. Adapted from M.Y. Sfeir et al. Science, 312, pp. 554–556, 2006, [150].
Adapted with permission from AAAS.
S S
of the (15,10) SWNT. As a result, a downshift of about 150 meV in the E33 and E44 transitions of
S S
the larger diameter nanotube is observed. The ratio of the E44 to E33 transition energies is similar
for these two nanotubes which are both of type I S-SWNT. However, a comparison of the (13,12)
type II S-SWNT with the (15,10) type I S-SWNT (Figure 14(b)) shows a different behavior for
the two different types of S-SWNTs. In this case, the average energies of the two transitions of
the (15,10) and (13,12) nanotubes are very similar due to their nearly identical diameters (1.71 nm
and 1.70 nm, respectively). However, the difference in behavior is manifested in the dissimilar
S S
intensity ratios for their E44 to E33 transitions. Rayleigh scattering measurements have been espe-
cially useful for advancing the theory related to optical transition energies in SWNTs and how
environmental effects influence these values [44]. Rayleigh scattering experiments have also been
carried out in graphene by Casiraghi [151].
Figure 15. (a) A 2D excitation vs. emission contour map for a dried (6, 5)-enriched DNA-CNT sample on a
sapphire substrate. The spectral intensity is plotted using the log scale shown on the right. (b) A schematic
view of the various observed light emissions plotted as the laser excitation energy vs. photon emission energy.
Reproduced figure with permission from S.G. Chou et al. Physical Review Letters 94, p. 127402, 2005
[154]. Copyright © (2005) by the American Physical Society.
emission spots, appearing in Figure 15(a) at E11 and denoted by circles, correspond to energies
where these bands cross the E11 transition energy. These intersection points are associated with a
mixture of PL and RRS processes involving one-phonon (VI) and two-phonon (I–V) processes,
following the labels in Figure 15(b). The various features in Figure 15 differ in linewidth. The
Raman peaks are much sharper (tens of cm−1 ) than the PL peaks which have linewidths of hun-
dreds of cm−1 . The various peaks also differ by the fact that, when changing the excitation laser
energy, the PL emission is fixed at E11 , while the Raman peaks change in frequency, keeping fixed
the energy shift from Elaser . With light emission occurring at the same energy, the PL and RRS
processes are sometimes confused in the literature, and the major reason is that RRS in solids often
has a much greater (typically 103 times larger) intensity than the non-resonance Raman spectra
[155]. To differentiate between the RRS and PL processes, one can just look at what happens to
the spectral output when changing Elaser .
The difference in linewidth between RRS and PL arises because in Raman scattering the
intermediate states that are excited between the initial state (incident photon plus the energy of
the system before light absorption) and the final state (emitted photon plus the energy of the
system after light emission) are “virtual” states.11 These virtual states do not have to correspond
to real states (and do not have to be eigenstates of the physical “system”) – any optical excitation
frequency will, in principle, suffice in RRS. In photoluminescence, on the other hand, the optically
excited state must be a real state of the system, and PL involves a real absorption of light at one
frequency, followed by a real emission of light at a different frequency.12 Photoluminescence is a
technique commonly used to study graphene samples engineered for an energy gap opening but
studies of energy gap opening are still at an early stage.
2.3.4. Electro-luminescence
Light emission by excited carriers can also be obtained by means other than photo-excitation, such
as electro-luminescence (EL). In EL the electrically induced excitation is followed by ambipolar
e–h recombination which comes from electrons and holes that are injected independently by using
doped semiconductor electrodes and impact excitation occurs by hot carriers. These effects have
444 R. Saito et al.
largely been studied in SWNTs [47,156,157], where both radiative decay of photo-excited and
electron-excited emission occur, as well as the non-radiative decay to create free carriers which can
then be studied by their photoconductivity spectra. EL processes could lead to the technological
use of carbon nanotubes as nanometer scale light sources and as photo-current or photo-voltage
detectors. EL has also been observed in graphene [157].
Figure 16. (a) Pump–probe time-delay data taken on a nanotube sample at a central wavelength of 800 nm
in (b). The decay of the pump–probe signal within several picoseconds reflects the decay of the excited-state
population. The inset in (a) is a zoom-in of the data between 0.3 and 6 ps, highlighting the coherent phonon con-
tribution to the signal. (b) CP oscillations excited and measured at five different Elaser excitations (expressed
in terms of their wavelengths). The individual traces for each wavelength are offset for clarity. The slower
decay of the excited-state population has been subtracted in each case. (c) Phonon spectrum detected at a
center wavelength of 765 nm obtained from both resonant Raman scattering (RRS) and from CP measure-
ments in the frequency range of the RBM [45]. Adapted from carbon Nanotubes: Advanced Topics in the
Synthesis, Structure, Properties and Applications, A. Hartschuh [45]. Copyright (2008) from Springer.
other crystal disorder, edge structure, strain, the number of graphene layers, nanotube diameter,
nanotube chirality and nanotube metallic vs. semiconductor behavior [169]. Another important
area where much work has been done is on disordered, amorphous and diamond-like carbons
[19,131,169], as well as graphite and graphene edges [151,170].
− +
Figure 17. ωG and ωG for semiconducting (filled circles) and metallic (open circles) SWNTs are plotted
+
as a function of 1/dt . The flat solid line shows ωG = 1591 cm−1 . The curves are given by the function
− −1
ωG = 1591 − C/dt , where C = CS = 47.7 cm nm for semiconducting SWNTs (long dashed curve) and
2 2
C = CM = 79.5 cm−1 nm2 for metallic SWNTs (short dashed curve). Also plotted (open squares) are the data
for the ∼ 1580 cm−1 Lorentzian peak sometimes observed in metallic SWNTs.Adapted with permission from
A. Jorio et al., Physical Review B 65, p. 75414, 2001 [171]. Copyright © (2001) by the American Physical
Society.
Advances in Physics 447
is called a micro-Raman set up since the spatial resolution of the Raman signal is of the order of
one micron. Not only is the incident light focused on the sample, but also the scattered light is
collected by the same optical lens and is split by a half mirror [173]. An isolated graphene flake
or SWNT on a Si substrate is located by putting the substrate on a mobile stage that moves the
sample horizontally, and the mobile stage is controlled by stepping motors.
Figure 18. Raman spectra imaging of an HOPG micro-crystallite. In (a) the G-band intensity is plotted. In
(b) the D-band intensity is plotted. (c) Spectra 1 and 2 are the spectra at locations 1 and 2 in (b) [174]. This
experiment was performed in the laboratory of Prof. Achim Hartschuh and represent, to our knowledge, the
first Raman imaging of localized defect modes in graphitic materials. M.A. Pimenta et al., Physical Chemistry
Chemical Physics 9, pp.1276–1290, 2007 [174]. Reprinted by permission of the PCCP Owner Societies.
Advances in Physics 449
occurs mainly at the edges of the crystalline region where symmetry-breaking occurs and D-band
intensity due to the edge discontinuity is also seen.
The intensity of the D-band can also be used to assign the atomic structure of the edge in
graphite and graphene [170], so that it can provide a useful tool to probe the edge chirality of
graphene. In particular, armchair edges have a large matrix element for D-band scattering while
for zigzag edges the matrix element for D-band scattering should vanish, and chiral edges show an
intermediate amount of D-band edge scattering. However, imperfect graphene edges produced by
the mechanical cleavage of graphite can produce ambiguous results that do not clearly discriminate
the armchair and zigzag edges from one another.
Figure 19. (a) RBM Raman measurements of HiPCO SWNTs dispersed in an SDS aqueous solution [152],
measured with 76 different laser lines Elaser [175]. The non-resonance Raman spectrum from a separated
CCl4 solution is acquired after each RBM measurement, and this spectrum is used to calibrate the spec-
tral intensities of each nanotube and to check its frequency calibration. (b) Filled circles are experimental
Eii vs. ωRBM data points obtained by Telg et al. [107] from analysis of an experiment very similar to
the one shown in (a). The label “Transition energy exp” actually indicates the excitation laser energy
(Elaser ) for each data point. Open circles come from third-neighbor tight-binding calculations, showing
that even the addition of interactions with more neighbors in the π-band based tight-binding model is
not enough to accurately describe the experimental results. Gray and black circles indicate the calculated
optical transition energies from semiconducting (E22 S and E S ) and from metallic (E M ) tubes, respec-
33 11
tively. (a) Adapted from M.J. O’Connell et al., Science, 297, p. 593, 2002 [152]. with permission from
AAAS. And adapted with permission from C. Fatini et al., Physical Review Letters 93, p. 147406, 2005
[175]. Copyright © (2005) by the American Physical Society. (b) Reprinted with permission from
H. Telg, et al., Physical Review Letters 93, p. 177401, 2004 [107]. Copyright © (2004) by the American
Physical Society.
ωRBM , and we call this plot an experimental Kataura plot. Figure 19(a) shows a 2D RBM map for
the HiPCO nanotube sample in aqueous solution wrapped by the SDS (sodium dodecyl sulfate)
surfactant [175]. For the construction of the plot in Figure 19(a), 76 different laser lines were
used. By fitting each of the spectra with Lorentzians, (n, m) indices were assigned to the different
SWNTs. Solid circles in Figure 19(b) denote the Eii values obtained experimentally by fitting the
resonance windows extracted from similar data to that shown in Figure 19(a), as compared with
the Eii obtained from tight-binding calculations (open circles) shown in Figure 19(b) [107].
near the metallic nano-particle is given by the boundary condition for the transmitted electromag-
netic wave on the metallic surface. When the light beam on the sample is located at the metallic
nano-particle, the corresponding Raman signal becomes significantly enhanced (sometimes up to
1010 times that of the normal Raman signal). The surface-enhanced Raman spectroscopy (SERS)
technique was utilized initially to obtain a large enough signal to observe the spectra from an
isolated individual SWNT [128,179–181]. Changes in the symmetry selection rules are observed
due to local symmetry breaking [182]. SERS measurements were carried out on SWNTs before
researchers realized that there was a strong resonance effect occurring in the π -related states in
carbon nanotubes and that this resonant effect was large enough to generate a measurable signal,
without any SERS enhancement effect. In fact the observation of strong SERS spectra stimulated
the first effort to observe the spectrum from one individual carbon nanotube [111]. Recently, the
use of the SERS technique has been applied to graphene [183–186].
Figure 20. Imaging of DNA-wrapped SWNTs at different magnifications: (a) A confocal Raman image of a
DNA-wrapped SWNT using an excitation wavelength of 632.8 nm. Topographic images in (b) and (c) indicate
a periodic height modulation expected for wrapping with short DNA segments. Near-field Raman images
(d) and (e) show the G-band intensity around 700 nm, and in (f) and (g) the intensity of photoluminescence
images at around 950 nm corresponding to the emission wavelength of an (8,3) nanotube are shown. PL
occurs only in the lower section of the nanotube where the Raman intensity is significantly weaker. The
abrupt transition from strong to weak Raman scattering combined with the appearance of PL is interpreted as
a local change in the nanotube (n,m) chirality [45]. Here, we see the power of near-field spectroscopy to show
images with spectroscopic information at high spatial resolution. Adapted from carbon Nanotubes: Advanced
Topics in the Synthesis, Structure, Properties and Applications, A. Hartschuh [45]. Copyright (2008) from
Springer.
Since CARS observes a blue-shifted signal, the signal is well separated from the related red-shifted
photoluminescence signal.
Figure 21. The dependence of the Raman G peak frequency of mono-layer graphene on doping using a gate
voltage to provide positive and negative potentials. (A) The G-band spectra at 295 K for many values of the
gate voltage Vg are shown. The darker line spectrum is at V = 0 but the spectrum at V = 0.6 V corresponds to
the undoped case, which occurs at V = 0 due to the natural doping of graphene by the environment. (B) The
G peak position (frequency) and (C) the G peak FWHM linewidth as a function of electron concentration
as deduced from the applied gate voltage data are shown. Black circles show the measurements and the
solid lines show results from a finite-temperature non-adiabatic calculation. Adapted with permission from
Macmillan Publishes Ltd. Nature Nanotechnology [195], Copyright © (2008).
Advances in Physics 455
Figure 22. On the left, the peak frequency (Pos(G)) and linewidth (FWHM(G)) for the Raman G-band
feature of doped bi-layer graphene vs. Fermi energy are shown. The black circles show the measurements
and the solid line shows the finite-temperature non-adiabatic calculation. On the right, schematics of the
el–ph coupling at three different doping levels, as indicated by the thicker lines on the electronic bands, are
shown. Adapted figure with permission from A. Das et al., Physical Review B 79, p. 155417, 2009 [191].
Copyright © (2009) by the American Physical Society.
456 R. Saito et al.
point phonons, which is an over-tone phonon mode and the overtone mode of the G-band appears
at around 3170 cm−1 (twice the G-band mode frequency at 1585 cm−1 ). A feature associated with
the harmonic of the D -band also appears at 3240 cm−1 .
The other process is given by emitting two K point phonons with q and −q wavevectors by
an inter-valley scattering process. By emitting two phonons with opposite q and −q vectors, the
photo-excited electron can go back to its original position and recombine with a hole. Because
of the many possible q = 0 vectors, the spectral width for two-phonon Raman processes is broad
compared with the linewidths typically found in the first-order Raman spectra. The G -band around
2700 cm−1 arises from a two-phonon Raman process occurring near the K point iTO phonon mode.
For both cases of two-phonon scattering processes, there is no restriction on q = 0 for either of
these processes. Furthermore, the two phonons are not always the same types of phonon modes
and can have the same wavevector but different frequencies. If two different phonon modes are
involved in the two-phonon process, the Raman shift becomes the sum of the two phonon mode
frequencies, which we call a combination mode. It is important to emphasize that combination
modes which combine via an intra-valley scattering process together with an inter-valley scattering
process are not possible since the corresponding q vectors would be completely different from
each other14 and conservation of momentum would be violated.
Figure 23. Schematic diagrams for (a1,a2) first-order and (b1–b4) one-phonon second-order, (c1,c2)
two-phonon second-order, resonance Raman spectral processes for which the top diagrams refer to inci-
dent photon resonance conditions and the bottom diagrams refer to the scattered resonance conditions. For
one-phonon, second-order transitions, one of the two scattering events is an elastic scattering event (dashed
line). Resonance points are shown as solid circles [104,207–209]. Adapted with permission from R. Saito et
al., New Journal of Physics 5, p. 157, 2003 [209]. Copyright © (2003) by the Institute of Physics.
intermediate state back to k (see Figure 23(b) and (c)) [209]. This two-scattering amplitude process
is expressed by perturbation theory in which the numerator of the resulting term consist of four
scattering matrix elements (two for photon absorption and emission, and two for phonon emission
or absorption), while the denominator of this term consists of three energy difference factors. If
two of the three energy difference factors becomes zero, the scattering intensity becomes strongly
enhanced by each of these factors. We call a process containing two resonance denominators
double resonance (DR) Raman scattering [104,174,207,210,211]. The G -band of mono-layer
graphene represents a resonance Raman spectral feature for an iTO phonon mode near the K point
which is resonant for each of the three scattering processes.
Figure 24. (a) The schematic diagram shows the light-induced e–h formation and the one electron–one
phonon scattering event taking place in the DR process with two different excitation laser energies (associated
with phonon wave vectors q1 and q2 , respectively), which are indicated by the gray and black arrows,
respectively. The two events in the DR process can occur in either order in time. (b) The phonon dispersion in
graphene is shown where the phonon wavevector q that fulfills the DR requirements for each Elaser value in
(a) is also indicated in terms of the phonon wavevectors q1 and q2 (see text). Reprinted figure with permission
from A. Jorio et al. Spectroscopy in Graphene Related Systems, 2010 [1]. Copyright © Wiley–VCH Verlag
GombH & Co. KGaA.
where K is the magnitude of the reciprocal lattice vector that connects K and K , and k (k ) here
is measured from the K (K ) point, which means that the double resonance wave vector qDR is the
Advances in Physics 459
Figure 25. The full DR Stokes Raman processes for inter-valley (a,b) and intra-valley (c,d) scattering. Here
(a,c) relates to the backward scattering process with qDR = k + k and (b,d) relates to the forward scattering
process with qDR = k − k , with k and k measured from the K point. The reciprocal lattice vector K is the
distance between the K and K points, k (k ) is the distance of the resonant states from K (K ), as defined in
(a). Reprinted figure with permission from A. Jorio et al. Spectroscopy in Graphene Related Systems, 2010
[1]. Copyright © Wiley–VCH Verlag GombH & Co. KGaA.
Figure 26. One of the possible DR Stokes Raman processes involving the emission of a phonon with wavevec-
tor −q. The set of all phonon wavevectors q which are related to transitions from points on the two circles
around K and K gives rise to the collection of small circles around the K point obeying the vector sum rule
q = K − k + k (here we neglect the trigonal warping effect). Note that this collection of circles is confined
to a region between the two circles with radii qDR = k + k ≈ 2k and qDR = k − k ≈ 0. The differences
between the radii of the circles around K and K and thus the radius of the inner circle around K are actually
small in magnitude and are here artificially enlarged for clarity in presenting the concepts of the double res-
onance process. Adapted with permission from L.G. Cançado et al., Physical Review B 66, p. 35415, 2002
[216]. Copyright © (2002) by the American Physical Society.
460 R. Saito et al.
phonon wavevector distance linking the K and K points. In the case of intra-valley scattering, we
just put K = 0 into Equations (11) and (12). Since the phonon energy is usually small compared
to the excited electronic energy levels, k ≈ k , these two double resonance conditions approach
qDR = 2k and qDR = 0 (as is commonly used in the literature [114,172,210]).
Elaser = 2vF k,
Eph = vph qDR , (13)
qDR = k ± k ,
where Elaser and Eph are, respectively, the laser and phonon energies, and k is the scattered electron
wave vector near the K point in the graphene BZ. It is important to remember that we are dealing
here with combination modes, so that the observed Eph has to reflect this combination. For example,
for the G -band, the observed G -band energy is given by EG = 2Eph , where Eph is the energy for
Advances in Physics 461
the iTO phonon mode at qDR .16 Making another commonly used approximation in Equation (13),
i.e., qDR = k + k ≈ 2k, then EG can be written as [1]
vph
EG = 2 Elaser . (14)
vF
A drawback in using the DR Raman features to define the electron and phonon dispersion relations
is that the measured values depend on both vph and vF , and one has to be known in order to obtain
the other. In addition to this problem, the physics of the phonon dispersion for graphene near
the K point is rather complex due to the Kohn anomaly, and the KA also occurs for phonons at
q → K (see Section 2.7). The high frequency of the iTO phonon when combined with the KA
near the K point are together responsible for the strong dispersive behavior observed for ωG . The
exact values for vph and vF are still under debate since they also depend on the complex physics
of many-body effects [112,194,217–220]. This is one area where more work for future research
is needed.
Figure 27. (a) Raman spectra of the G and the G∗ -bands of mono-layer graphene for 1.92, 2.18, 2.41, 2.54
and 2.71 eV laser excitation energy. (b) Dependence of ωG and ωG∗ on Elaser . The circles correspond to
the graphene data and the lozenges correspond to data for turbostratic graphite. Adapted with permission
from D. Mafra et al., Physical Review B, 76, p. 233407, 2007 [213]. Copyright © (2007) by the American
Physical Society.
462 R. Saito et al.
frequencies ωG and ωG∗ as a function of Elaser for graphene and turbostratic graphite (for which
the stacking between graphene layers is random). The G -band in Figure 27 exhibits a highly
dispersive behavior with (∂ωG /∂Elaser ) 88 cm−1 /eV for mono-layer graphene, 95 cm−1 /eV for
turbostratic graphite [213] and 106 cm−1 /eV for carbon nanotubes (see Figure 27 and [222]). The
G∗ -band feature exhibits a much less pronounced dispersion than the G -band, and of opposite
sign, with (∂ωG∗ /∂Elaser ) −10 to −20 cm−1 /eV for both mono-layer and turbostratic graphite
[208,213], and no dispersion is reported for carbon nanotubes [221]. It should be noted that a
different interpretation to the origin of the G∗ -band is given in [131,223], which together with
[221] identified the origin of the 2450 cm−1 peak with the overtone of the 1225 cm−1 feature
which has a peak in the phonon density of states for two phonons [223], while [224,225] assigned
this feature to the combination modes iTA + iTO. As already stated, the qDR ≈ 2k wavevector
gives rise to the G -band, while the qDR ≈ 0 wavevector gives rise to a DR feature coming from
the iTO phonon very close to the K point. The qDR ≈ 0 processes are expected to be less intense
than the qDR ≈ 2k processes because the destructive interference condition is exactly satisfied for
qDR = 0 [208].
2.9. Summary
The power of Raman spectroscopy for studying carbon nanotubes is in particular revealed through
exploitation of the resonance Raman effect, which is greatly enhanced by the singular density
of electronic states of SWNTs and the resonant effect comes from the 1D confinement of the
electronic states due to the small diameters of carbon nanotubes. Soon after the discovery of the
Advances in Physics 463
Figure 28. Raman spectra of graphite whiskers obtained at three different laser wavelengths (excitation
energies) [226]. Note that some phonon frequencies vary with Elaser and some do not. Above 1650 cm−1 the
observed Raman features are all multiple-order combination modes and overtones (see Figure 29), though
some of the peaks observed below 1650 cm−1 are actually one-phonon bands activated by defects. The inset to
(c) shows details of the peaks labeled by L1 and L2 . The L1 and L2 peaks, which are dispersive, are explained
theoretically by the defect activation of double-resonance one-phonon processes (see Section 2.8.10) involv-
ing the acoustic iTA and LA branches, respectively, as discussed in Ref. [210]. Adapted figure with permission
from P.H. Tan et al., Physical Review B 64, p. 214301, 2001 [226]. Copyright © (2001) by the American
Physical Society.
resonance Raman effect in SWNTs [103], it was found that the resonance lineshape could be used
to identify the nanotube structure, i.e., the chiral indices (n, m) [111], and to distinguish metallic
from semiconducting SWNTs [128,228]. It is clear that most of the results achieved up to now
have been developed for SWNTs, while the optics of graphene and nanoribbons is still at an early
stage. This is the way it happened historically, and the knowledge developed in carbon nanotube
science is now fostering an amazingly fast development of graphene photo-physics. It is expected
that graphene photo-physics will follow a similar path of development that will reveal much new
physics as this very fundamental field develops such as the understanding of the KA which helped
to elucidate the phonon dispersion relation of graphite and all related sp2 carbons.
464 R. Saito et al.
Figure 29. Two-phonon dispersion of graphite based on second-order double-resonance peaks in the Raman
spectra (dark circles). Solid lines are dispersion curves from ab initio calculations considering combination
modes and overtones associated with totally symmetric irreducible representations. Adapted figure with
permission from J. Maultzsch et al., Physical Review B 70, p. 155403, 2004 [208]. Copyright © (2004) by
the American Physical Society.
within the resonance window, so care must be exercised in making proper comparisons between
experiment and theory.
3.1.5. Excitons
In the case of carbon nanotubes, the exciton binding energy is much larger (up to 1 eV)
[120–122,147,242] than that for Si (which is in the meV range). The exciton, which is
formed from a photo-excited electron and the hole that is left behind, is especially impor-
tant and dominates the observed optical processes in carbon nanotubes which are 1D systems
where excitonic effects are exceedingly strong. Even at room temperature, the excitoni-
cally mixed electronic states are specified by a wavevector k so as to form a spatially
localized state. In order to obtain excitonic states and their corresponding wavefunctions,
the Bethe–Salpeter equation for π electrons is used here within the tight-binding method
(Section 3.5). Using excitonic wavefunctions, we can calculate the relevant exciton–photon
(Section 3.6) and exciton–phonon (Section 3.6.3) matrix elements. Two-photon absorption or
time-dependent Raman spectroscopy have also been used to observe many specific exciton-related
phenomena.
where Mi and ui are, respectively, the mass and the vibrational amplitude of the ith atom and K (ij)
represents a 3 × 3 force constant tensor which connects ith and jth atoms. The summation on j is
taken over the jth nearest neighbor atoms so as to reproduce the phonon energy dispersion relation
(see Figure 30). The K (ij) terms are obtained by fitting to experimentally obtained phonon disper-
sion relations, such as are determined from neutron or X-ray inelastic scattering measurements
[208,235,239]. The fitting procedure to the experimental phonon dispersion is possible even if the
KA effect is included. However, the broadening of the phonon dispersion due to the finite lifetime
Advances in Physics 467
Figure 30. Phonon dispersion of graphene in the 2D BZ. The symbols are experimental data obtained by
inelastic X-ray scattering [244] and the lines are fitted to the experimental phonon data using up to 14th
nearest-neighbor interactions [245].
of phonon cannot be expressed by the present method and the self-energy for the phonon should
then be calculated as discussed below [196,205,219,220,246].
Since the lattice is periodic, each displacement ui in the unit cell can be expressed by a wave
with a phonon wavevector k and frequency ω as follows:
1 i(k·Ri −ωt)
uk(i) = √ e ui , (16)
Nu R
i
where the sum is taken over all Nu lattice vectors Ri in the crystal for the ith atom in the unit cell.
The equation for uk(i) (i = 1, . . . , N), where N is the number of atoms in the unit cell, is given by
[32]
⎡ ⎤
⎣ K (ij) − Mi ω2 (k)I ⎦ uk(i) −
(j)
K (ij) eik·Rij uk = 0 (17)
j j
in which I is a 3 × 3 unit matrix and Rij = Ri − Rj denotes the relative coordinate of the ith
atom with respect to the jth atom. The simultaneous equations implied by Equation (17) with 3N
unknown variables uk ≡ t (uk(1) , uk(2) , . . . , u(N)
k ), for a given k vector, can be solved by a diagonal-
ization of the 3N × 3N matrix in brackets, which we call the dynamical matrix. By diagonalizing
the dynamical matrix for each k, we get the phonon frequencies and corresponding amplitudes
as a function of k, ω(k) and uk , respectively, which are the eigenvalues and eigenfunctions of the
dynamical matrix.
In Figure 30, the phonon dispersion relations of graphene are plotted in the 2D BZ. Lines
are fitted for the calculated phonon energy dispersion to the experimental data for inelastic X-ray
scattering (symbols) by a set of force constants that includes force constants up to 14th nearest
neighbors [245]. This force constant set is obtained by minimizing the square of the difference
between experiment and theory for each experimental data point. In order to get good convergence
for the nonlinear fitting, we must start with a small number of nearest neighbors and we then
increase the number of neighbors one by one. Further, in order to get the required zero value for
the acoustic phonon modes at the point, we should consider the relationships between the force
constant set, known as the force constant set sum rule [217]. Degenerate in-plane optical phonon
modes around 1600 cm−1 at the point are known by symmetry requirements to correspond to the
Raman-active phonon mode (G-band), while the out-of-plane optical (oTO) phonon mode around
860 cm−1 at the point is an infrared-active phonon mode. The acoustic modes are discussed
in [247].
468 R. Saito et al.
The phonon modes near the K point and point can be observed by defect-induced or
two-phonon derived features in the Raman spectra. The phonon modes along the phonon dis-
persion relation can be observed by studying phonon modes arising from DR Raman processes
(see Section 2.8.6) [210]. The LO phonon mode (the highest frequency mode) of graphene has
a local minimum at the point and the phonon energy increases with increasing k by a pro-
cess which we call “over-bending”. The “over-bending” can be reproduced using force constants
going beyond the fifth nearest neighbor. Both the in-plane optic phonon modes near the point
and the iTO phonon mode near the K point show phonon softening phenomena for graphene and
metallic SWNTs and the resulting phonon frequency down-shifts are known as the KA effect
[127,194,196–199,205,219,220,243,246] (see Section 3.6.5). When we calculate phonon disper-
sion by first principles [248–250], the effect of the Kohn anomalies should be taken into account
when calculating a force constant set, while in the simple tight-binding method, we just obtain
a force constant set either by fitting to the experimental results [208,244,251,252] or by first-
principles calculation in which the Kohn anomalies are taken into account. When we obtain a
force constant set by the atomic potential as a function of the C–C bond distance, such as by the
extended tight-binding method [253,254], we should consider the additional effect of the KA in
the calculation.
Interlayer force constants of multi-layer graphene are much weaker than the intralayer force
constant set. Each phonon mode of a graphene sheet is split into symmetric and anti-symmetric
vibrational modes with respect to the inversion or mirror symmetry of multi-layers, depending on
whether the multilayer graphene consists of an even or odd number of graphene layers, respectively.
It is important that some phonon modes (oTO, oTA, LO) become Raman active (or inactive) by
the interlayer interaction [4].
N
j (k,
r) =
Cjj (k)
j (k,
r) (j = 1, . . . , N), (18)
j =1
are coefficients to be determined and N the number of atomic orbitals in the unit
where Cjj (k)
cell. When we consider π orbitals for n-layer graphene, then N = 2n. Here j denotes the Bloch
function for an atomic orbital ϕj which is given by
1
Nu
j ( k,
r ) = √ )
eik·R ϕj (r − R (j = 1, . . . , N), (19)
Nu
R
where the summation takes place over Nu lattice vectors R in the crystals. When we put
Equation (18) into the Schrödinger equation H j = E j for a Hamiltonian H, we get
N
N
ij = Ei (k)
Hjj (k)C ij
Sjj (k)C (i = 1, . . . , N). (20)
j =1 j =1
Advances in Physics 469
=
Hjj (k) j |H| j ,
=
Sjj (k) j| j (j, j = 1, . . . , N). (21)
HCi = Ei (k)SCi. (23)
By diagonalization of a given H and S for each k vector, we get the energy eigenvalues Ei (k)
and
eigenfunctions Ci (k).
The ij matrix element of H is given by
= 1 ik(R−R )
Hij (k) e ϕi (r − R )|H|ϕj (r − R)
Nu R,R
(24)
= eik(R) ϕi (r − R)|H|ϕj (r),
R
where R ≡ R − R and in the second of line of Equation (24), we use the fact that the tight-
binding parameter ϕi (r − R )|H|ϕj (r − R) only depends on R. Similarly, the matrix elements
of S are given by
=
Sij (k) eik(R) ϕi (r − R)|ϕj (r). (25)
R
The tight-binding parameters ϕi (r − R)|H|ϕj (r) and ϕi (r − R)|ϕj (r) are obtained by:
(1) integrating the matrix elements using the atomic orbitals ϕj (r) [247] or (2) fitting them so as
to reproduce experimentally obtained energy dispersion measurements. Values for a typical fitted
parameter set (TBP) are listed in Table 1.
As seen in Table 1, tight-binding parameters are listed for up to the third nearest neighbor
(3NN) within a layer (upper half) and for interlayer interactions between graphene layers (lower
half) [255]. In Figure 31, we show a definition of the tight-binding parameters listed in Table 1 for
the Hamiltonian for pairs of carbon atoms separated by their corresponding distances R [255].
j
The notation used for the parameters γi follows that of Slonczewski and Weiss [257], while γ0 and
sj (j = 1, 2, 3) denote the in-plane parameters with the jth nearest neighbors up to the 3rd nearest
neighbor (3NN). As far as we consider transport properties near the K point of the first BZ, the
in-plane nearest-neighbor parameter γ01 is sufficient. However, when we consider optical transition
phenomena around the K point, it is necessary to include the parameters γ02 and γ03 which are
indicated explicitly in Figure 31 [256]. The parameters γ1 , γ3 and γ4 denote interactions between
carbon atoms in the nearest-neighbor layers, while the parameters γ2 and γ5 couple carbon atoms
in next nearest-neighbor layers. The parameters γ3 and γ4 introduce a k-dependent interlayer
interaction and γ2 sensitively determines a small energy dispersion along the KH direction in the
3D BZ for energy bands near the Fermi energy of graphite (see [5, Fig. 1f]) which gives rise to
the semi-metallic nature of 3D graphite.
The overlap tight-binding parameters, s0 , s1 and s2 , are essential for describing the asymmetry
between the valence and conduction energy bands relative to the Fermi energy. The energy band
470 R. Saito et al.
TBP 3NN TB-GWa 3NN TB-LDAa EXPb 3NN TB-LDAc R, paird
√
γ01 −3.4416 −3.0121 −5.13 −2.79 a/ 3, AB
γ02 −0.7544 −0.6346 1.70 −0.68 a, AA and BB
√
γ03 −0.4246 −0.3628 −0.418 −0.30 2a/ 3, AB
√
s0 0.2671 0.2499 −0.148 0.30 a/ 3, AB
s1 0.0494 0.0390 −0.0948 0.046 a, AA and BB
√
s2 0.0345 0.0322 0.0743 0.039 2a/ 3, AB
γ1 0.3513 0.3077 – – c, AA
γ2 −0.0105 −0.0077 – – 2c, BB
√
γ3 0.2973 0.2583 – – (a/ 3,c), BB
√
γ4 0.1954 0.1735 – – (a/ 3,c), AA
γ5 0.0187 0.0147 – – 2c, AA
E0e −2.2624 −1.9037 – −2.03
f 0.0540g 0.0214 – –
a
Fits to LDA and GW calculations [255].
b
Fit to ARPES experiments by Rotenberg et al. [124].
c
Fit to LDA calculations by Reich et al. [256].
d
In-plane and out-of-plane distances between a pair of A and B atoms.
e
The energy position of π orbitals relative to the vacuum level.
f
Difference of the diagonal term between A and B atoms for multi-layer graphene.
g
The impurity doping level is adjusted in order to reproduce the experimental value of in graphite.
All values are in eV except the dimensionless overlap parameters of s0 -s2 . The parameters of fits to LDA and
GW calculations are shown. The 3NN Hamiltonian is valid over the whole 2-D (3-D) BZ of graphite (graphene
layers) [255]. Mopac93 and Gaussian 9 software packages were used for implementing Gaussian and other
software applications.
Figure 31. Identification of the various Slonczewski–Weiss parameters for the tight-binding parameters for
a pair of carbon atoms separated by a distance R. Adapted figure with permission from A. Grüneis et al.,
Physical Review B 78, p. 205425, 2008 [255]. Copyright © (2008) by the American Physical Society.
width of the conduction band is larger than that of the valence band when using this set of tight-
binding parameters [32], thereby inducing asymmetry between the electrons and holes in few layer
graphene. Further, depending on whether the number of graphene layers is odd (even), the linear
(quadratic) k energy dispersion behavior appears near the K point. Koshino and Ando [258] have
explained analytically the reason for the odd-even dependence of the electronic structure of few
layer graphene on the number of graphene layers by the tight-binding method.
Advances in Physics 471
where P is the unit vector (polarization vector) which specifies the direction of E, I the intensity
of the light in W/m2 and 0 the dielectric constant for the vacuum in SI units. The “±” sign
corresponds to the emission (“+”) or absorption (“−”) of a photon with frequency ω. Here, we
can assume that the wavevector k of an electron does not change during the transition (vertical
transition). Then the matrix element for optical transitions from an initial state ı (k) to a final
472 R. Saito et al.
f
state (k) at k is defined by
fı
Mopt (k) = f
(k)|Hopt | ı
(k). (28)
The electron–photon matrix element between initial and final states in Equation (28) is calculated
by
fı e Iρ i(ωF −ωı ±ω)t f ı
Mopt (k) = e D (k) · P (29)
mωρ c0
where the weak spatial dependence of the vector potential A is neglected and Df ı (k) is the dipole
vector defined by the matrix element
Df ı (k) = f
(k)|∇| ı
(k). (30)
For a given polarization, P, the optical absorption (or stimulated emission) becomes large (absent)
when D is parallel (perpendicular) to P.
f∗
Df ı (kF , kı ) = CB (kF )CAı (kı ) B (kF , r)|∇| A (kı , r)
(31)
f∗
+ CA (kF )CBı (kı ) A (kF , r)|∇| B (kı , r).
Since both the 2pz orbital and the ∂/∂z component of ∇ have odd symmetry with respect to the z
mirror plane, the z component of D becomes zero. Thus, we conclude that the dipole vector lies
in the xy plane.
When we expand into atomic orbitals, the leading term of A (kF , r)|∇| B (kı , r) is the
atomic matrix element mopt between nearest neighbor atoms given by
∂
mopt
= φ(r − Rnn ) φ(r) , (32)
∂x
where Rnn is the lattice vector between nearest-neighbor C atoms along the x-axis.
When we use a linear approximation for the coefficients CA and CB for a k point around the
corner point of the 2D BZ K = (0, −4π/(3a)) for the valence (v) and conduction (c) bands, we
Advances in Physics 473
Figure 32. (a) The normalized dipole vector Dcv (k) is plotted as a function of k over the 2D BZ. (b)
The oscillator strength in units of the atomic matrix element mopt is plotted as a function of k over
the 2D BZ. The separation between two adjacent contour lines is 0.4 mopt . The darker areas have a
larger value for the oscillator strength. Reprinted figure with permission from A. Grüneis et al., Physical
Review B 67, pp. 165402–165407, 2003 [262]. Copyright © (2003) by the American Physical Society.
write [32]
1 ky − ikx
CAv (K + k) = √ , CBv (K + k) = √ .
2 2k
(33)
1 −ky + ikx
CAc (K + k) = √ , CBc (K + k) = √ ,
2 2k
The electric dipole vector coupling the valence and conduction bands is then given by
3mopt
Dcv (K + k) ≡ c
(K + k)|∇| v
(K + k) = (ky , −kx , 0). (34)
2k
In Figure 32(a) we plot the normalized directions of the normalized dipole vector Dcv (k) as arrows
over the 2D BZ of graphene [262]. Around the K points, the arrows show a vortex behavior. Note
also that the rotational directions of Dcv (k) around the K and K points are opposite to each other
in Figure 32(a). In Figure 32(b) we plot the values of the magnitude of the oscillator strength O(k)
in units of mopt on a contour plot. Here Ocv (k) is defined by
Ocv (k) = Dcv∗ (k) · Dcv (k). (35)
As shown in Figure 32(b), the oscillator strength Ocv (k) has a maximum at the M points and a
minimum at the point in the 2D BZ. The k dependent Ocv (k) will be relevant to the calculation for
the type-dependent photoluminescence (PL) intensity of a single wall S-SWNT [263] in which the
PL of type I (mod(2n + m, 3) = 1) is stronger than for type II (mod(2n + m, 3) = 2) S-SWNTs,
though we need to consider the electric dipole vector for each carbon nanotube individually in
terms of its diameter and chiral angle [262,264,265].
The optical absorption intensity is given by the inner product Dcv (k) · P up to linear terms in
kx and ky for a given polarization vector P = (px , py , pz )
3mopt
P · Dcv (K + k) = ± (py kx − px ky ). (36)
2k
Equation (36) shows that the line py kx − px ky = 0 in the 2D BZ denotes the conditions for the
occurrence of a node in the optical absorption for a given polarization vector P = (px , py ). In
the case of graphene, however, the optical transition events take place along equi-energy contours
474 R. Saito et al.
around the K points, and we cannot see the nodes. This phenomena might be observed in graphene
nanoribbons in which a 1D k value is specially selected. The polarization dependence of the optical
absorption relative to the edge of graphene nanoribbon is now an interesting problem. In the case
of a normal semiconductor, since the dipole vector is not a linear function of k, we cannot get a
node in such cases.
where s = A and B is an index denoting the two carbon atoms in the unit cell, and Rt denotes the
equilibrium atom positions relative to the origin. φt,o denotes the atomic wave functions for the
orbitals o = 2s, 2px , 2py and 2pz at Rt , which are real functions (with no imaginary components).
The potential energy of the lattice V can be expressed by the atomic potentials v(r − Rt ) at
Rt ,
V= v(r − Rt ), (38)
Rt
where v in Equation (38) is given by the first-principles calculation for the Kohn–Sham potential
of a neutral pseudo-atom [253]. The matrix element for the potential energy between the two
different states i = a,k and f = a ,k is then written as
1 ∗
a ,k (r)|V | a,k (r) = C (a , k )Cs,o (a, k)
Nu s ,o s,o s ,o
× ei(−k ·Ru ,s +k·Ru,s ) m(t , o , t, o), (39)
u u
The atomic matrix element m thus comes from an integration over three centers of atoms, Rt ,
Rt and Rt . We neglect m for the cases for which the three centers t, t and t are different from
one another. When we consider only two center integrals, m consists, respectively, of off-site and
Advances in Physics 475
The potential v(r − Ri ) is vibrating within an adiabatic approximation with a phonon amplitude
S(Rt ). Then the potential modification δV due to a lattice vibration is given by
δV = v[r − Rt − S(Rt )] − v(r − Rt )
Rt
(42)
≈− ∇v(r − Rt ) · S(Rt ).
Rt
Since the potential modification δV breaks the periodicity of the lattice, the wavevector for an
electron is no longer a good quantum number and thus the scattering of an electron by the el–ph
interaction occurs. If we consider δV as a perturbation, then the el–ph matrix element is defined
on the basis of perturbation theory: as [139,267–270]
Ma,k→a ,k ≡ a ,k (r)|δV |
a,k (r)
1 (43)
=− Cs∗ ,o (a , k )Cs,o (a, k) ei(−k ·Ru ,s +k·Ru,s ) δm(t , o , t, o),
Nu s ,o s,o u ,u
where δm(t , o , t, o) is the atomic deformation potential which consists of the off-site and on-site
deformation potentials δmα and δmλ given by
δmα = φs ,o (r − Rt ){∇v(r − Rt ) · S(Rt ) + ∇v(r − Rt ) · S(Rt )} × φs,o (r − Rt ) dr,
⎧ ⎫
⎨ ⎬
δmλ = δRt ,Rt φs ,o (r − Rt ) ∇v(r − Rt ) · S(Rt ) φs ,o (r − Rt ) dr. (44)
⎩ ⎭
Rt =Rt
It is noted that both terms δmα and δmλ are of the same order of magnitude and that they work in
a different way for each phonon mode [271].
The atomic deformation potential for any orbitals and for any vibration can be expressed by a
small number of terms which are defined by the bonding or force constants between atoms along
or perpendicular to the two atoms by using the Slater–Koster scheme [32,253]. The atomic el–ph
matrix elements φ|∇v|φ are thus defined for four fundamental hopping and overlap integrals
denoted by (ss), (sσ ), (σ σ ) and (ππ), which are defined as a function of the C–C distance [253,
268] as follows:
α p (τ ) = φμ (r)∇v(r)φν (r − τ) dr = αp (τ )Î(αp ),
(45)
λ p (τ ) = φμ (r)∇v(r − τ)φν (r) dr = λp (τ )Î(λp ),
which are, respectively, denoted by α p (τ ) and λ p (τ ) for the off-site and on-site deformation
potentials. Here Î(αp ) and Î(λp ) are unit vectors describing the direction of the off-site and on-site
476 R. Saito et al.
Figure 33. (a) The nine non-zero off-site deformation potential vectors α p . The dashed curves represent
the atomic potentials. (b) The six non-zero on-site deformation potential vectors λ p . The dashed curves
represent the atomic potentials. For λss , λσ σ and λπ π , the two same orbitals are illustrated by shifting
them with respect to each other. Reprinted figure with permission from J. Jiang et al., Physical Review B
72, pp. 235408–235411, 2005 [268]. Copyright © (2005) by the American Physical Society.
Figure 34. (a) The off-site deformation potential α p and (b) the on-site deformation potential λ p as
a function of inter-atomic distance. The vertical line corresponds to 1.42 Å which is the C–C dis-
tance in graphite Reprinted figure with permission from J. Jiang et al., Physical Review B 72, pp. 235408–
235411, 2005 [268]. Copyright © (2005) by the American Physical Society.
deformation potential vectors α p and λ p , respectively [271], and p = μν as given in Equation (45).
The 2p orbital φμ (φν ) is along or perpendicular to the bond connecting the two carbon atoms and
τ is the distance between the two atoms.17 In Figure 33, we show the non-zero matrix elements for
p atomic deformation potentials for 2s, σ and π atomic orbitals.
the (a) off-site α p and (b) on-site λ
In Figure 34, the calculated values of α p and λ p are plotted as a function of inter-atomic
distance between two carbon atoms [268]. At r = 1.42 Å, which is the bond length between a
carbon atom and one of its nearest neighbors, we have απ π ≈ 3.2 eV/Å, and |λπ π | ≈ 7.8 eV/Å, and
|απ σ | ≈ 24.9 eV/Å. In order to calculate the el–ph matrix element of Equation (43) for each phonon
mode, the amplitude of the atomic vibration S(Rt ) for the phonon mode (ν, q) is calculated by
S(Rt ) = Aν (q) n̄ν (q)eν (Rt )e±iων (q)t . (46)
Advances in Physics 477
Here ± is for phonon emission (+) and absorption (−), respectively, and A, n̄, e and ω are the
zero-point phonon amplitude, number, eigenvector, and frequency, respectively. At equilibrium,
the phonon number in Equation (46) is determined by the Bose–Einstein distribution function
nν (q) for phonon ν:
1
nν (q) = ω/k T . (47)
e B −1
Here, T = 300 K is the lattice temperature at room temperature and kB is the Boltzmann constant.
For phonon emission, the phonon number is n̄ = n + 1, while for phonon absorption, n̄ = n. The
amplitude of the zero-point phonon vibration is
Aν (q) = , (48)
Nu MC ων (q)
where MC is the mass of a carbon atom and the phonon eigenvector eν (Rt ) is given by diagonalizing
the dynamical matrix Eq. (17)18 .
where E(kc ) and E(kv ) are the quasi-electron and quasi-hole energies, respectively (see
Equation (52)). Here “quasi-particle” means that the particle has a finite lifetime in an excited
state because of the Coulomb interaction. Equation (49) is solved by a matrix that includes many
478 R. Saito et al.
kc and kv points. The mixing term of Equation (49) which we call the kernel, K(kc kv , kc kv ), is
given by
K(kc kv , kc kv ) = −K d (kc kv , kc kv ) + 2δS K x (kc kv , kc kv ) (50)
with δS = 1 for spin singlet states and 0 for spin triplet states [167]. The direct and exchange
interaction kernels K d and K x are, respectively, given by [274]
where the functions w and v are the screened and bare Coulomb potentials, respectively, and ψ
denotes the quasi-particle wavefunction. The quasi-particle energies are the sum of the single
particle energy ((k)) and the self-energy ((k)),
In order to determine the kernel and self-energy, the single particle Bloch wavefunction ψk (r)
and the screening potential W are evaluated by either a first-principles calculation [147] or by using
an extended tight-binding wavefunction within a random phase approximation (RPA) calculation
[148]. In the RPA, the static screened Coulomb interaction for π electrons is expressed by
V
W= , (54)
κ(q)
with a static dielectric constant κ and a dielectric function (q) = 1 + v(q)(q). For describing
the exciton energy and exciton wavefunction it is essential to select a reasonable function for the
unscreened Coulomb potential v(q) [120,148]. For 1D materials, the Ohno potential is commonly
used for the unscreened Coulomb potential v(q) for π orbitals [275] at two sites, Ru s and R0s , (u:
unit cell, s: atom position) with
U
v(|Ru s − R0s |) = , (55)
((4π0 /e2 )U|Rus − R0s |)2 + 1
where U is the energy cost to place two electrons on a single site (|Rus − Ros | = 0) and this energy
cost is taken as U ≡ Uπa πa πa πa = 11.3 eV for π orbitals [275]. The Ohno potential works well in
reproducing the ground state and low-energy electronic excitations [276].
Advances in Physics 479
Mex-op = 0 |Hel-op |0
n
= n∗
Dk Zkc,kv , (56)
k
where 0n is the exciton wavefunction with a q = 0 center of mass momentum. Since the center
of mass momentum is conserved before and after an optical transition, only q = 0 excitons can
be excited.
In the case of a SWNT, since the lattice structure is symmetric under a C2 rotation around an
axis which is perpendicular to the nanotube axis and goes through the center of a C–C bond, the
C2 exchange operation between A and B carbon atoms in the hexagonal lattice is equivalent to
the exchange of k and −k states. Since the exciton wavefunction of a carbon nanotube should
transform as an irreducible representation of the C2 symmetry operation, we can get A1 , A2 , E
and E ∗ symmetry excitons [142]. For example, the A1 and A2 exciton wavefunctions which are,
respectively, symmetric and antisymmetric under a C2 rotation, are given by
1 n + +
| n
0 (A1,2 ) =√ Zkc,kv (ckc ckv ∓ c−kc c−kv )|0, (57)
2 k
where k and −k are located around the K and K points, respectively, and −(+) in ∓ corresponds
to an A1 (A2 ) exciton.19 When we use the relation Dk = D−k , the excitonic-optical (ex-op) matrix
elements for the A1 and A2 excitons are given by
Mex-op (An1 ) = 0,
√ (58)
Mex-op (An2 ) = 2 n∗
Dk Zkc,kv .
k
Equation (58) directly indicates that A1 excitons are dark and only A2 excitons are bright, which is
consistent with the predictions by group theory [277]. Because of the spatially localized exciton
wavefunction, the exciton–photon matrix elements are greatly enhanced (on the order of 100
times) compared with the corresponding electron–photon matrix elements [273].
+
where M(c) (M(v)) denotes the el–ph matrix element for the conduction (valence) band, and bqν
(bqν ) is a phonon creation (annihilation) operator for the νth phonon mode at q. From Equation (59),
we obtain the exciton–phonon matrix element between the initial state | q1 n1
and a final state | q2
n2
,
480 R. Saito et al.
by writing
Mex−ph = q2
n2
|Hel-ph | q1
n1
= [Mk,k+q
ν n2∗
(c)Zk+q,k−q1 n1
Zk,k−q1 − Mk,k+q
ν n2∗
(v)Zk+q2,k n1
Zk+q2,k+q ], (60)
k
where γ is the width of the resonance Raman window (Section 3.6.3) [278]. The γ value is
essential in determining Iel as a function of laser excitation energy (Raman excitation profile).
When we use the exciton–photon and exciton–phonon interactions, we apply the formula to the
Raman intensity of SWNTs as follows:
2
1 M
ex-op (a)Mex-ph (a → b)Mex-op (b)
Iex =
L (E − Ea + iγ )(E − Ea − Eph + iγ )
a
2 (62)
1 Mex-op (a)2 Mex-ph (a → a)
= .
L (E − Ea + iγ )(E − Ea − Eph + iγ )
a
In the second line of Equation (62), we assume that the virtual state b can be approximated by
the real state a.20 In the case of a first-order Raman process, since q = 0, the matrix element of
Equation (60) is simplified as
Mex-ph = [Mk,k
ν
(c) − Mk,k
ν
(v)]|Zk,k |2 . (63)
k
When we consider the second-order Raman intensity, we should consider q = 0 phonon scat-
tering. In this case, the exciton–phonon interaction between an A2 exciton state and an E exciton
state is important, in particular, for the case where the E exciton state consists of an electron near
the K point and a hole near the K point and vice versa. Here the inter-valley exciton–phonon
interaction is generally large.
in which the summation is taken over two intermediate electronic states m and m and the cor-
responding phonon frequencies ω1 and ω2 with phonon wavevectors −q1 and −q2 , respectively,
and for the initial states i, after taking the square of the scattering amplitude, Jm,m that is given by
where Em i ≡ Elaser − (Em − Ei ) and Mex-op (mi) denote the optical transition from i to m states,
etc. In general, energy and momentum conservation for the incident (i) and scattered (s) electrons
requires:
where + (−) in Equation (66) and − (+) in Equation (67) correspond to phonon absorption and
emission with the wavevectors q1 and q2 . By considering ks ≈ ki (see Section 1.5), momen-
tum conservation requires q2 ≈ −q1 for satisfying the DR condition for two of the three energy
denominators in Equation (65).
1
= Wkν
τν
S |Dν (k, k )|2 dE(μ , k ) −1
= (68)
8π Mdt μ ,k ων (k − k) dk
δ(ω(k ) − ω(k) − ων (k − k)) δ(ω(k ) − ω(k) + ων (k − k))
× + ,
eβων (k −k) − 1 1 − e−βων (k −k)
where S, M, dt , β and μ , respectively, denote the area of the graphene unit cell, the mass of a
carbon atom, the diameter of a SWNT, 1/kB T (where kB is the Boltzmann constant), and the
cutting line indices of the final state. Here Dν (k, k ) denotes a matrix for scattering an electron
from k to k by the νth phonon mode. The relaxation process is restricted to 24 possible phonon
scattering processes satisfying energy–momentum conservation [267]. The two terms in braces
in Equation (68), respectively, represent the absorption and emission processes of the νth phonon
482 R. Saito et al.
mode with energy ων (k − k). The calculation of the transition rates as in Equation (68) have
been considered by the Ferrari group using another approach [151,219].
For the result, in the case of S-SWNTs, we can obtain calculated γ values in agreement with
experiments by just considering the electron–phonon coupling model [278]. The calculated γ value
shows a strong dependence on chirality and diameter for S-SWNTs. However, the calculated γ
value for M-SWNTs is much smaller than the experimental γ value which shows the presence of
Figure 35. (a) The calculated G-band spectra for S-SWNTs with the same family number p = 2n + m = 28.
(b) The calculated electron–phonon matrix elements vs. chiral angle θ for the LO and TO phonons and
for two different 2n + m family numbers (22 and 28). (c) Plot of γ vs. θ for members of p = 28. (d)
Plot of R vs. θ for three 2n + m families of M-SWNTs. (e) The angle R between the circumferential
vector K1 and the cutting line for the polar coordinate of a k vector at the van Hove singular point. (f)
The angle φ between the tube axis and the phonon eigenvector direction for a (12,6) SWNT. The calcu-
lated angles φ vs. θ for the TO phonons (g) and the LO phonons. (h) For the results for the LO phonons
as a function of θ (fitted by the function of Equation (70) (see text)). Reprinted with permission from
J.S. Park et al., Physical Review B 80, p. 81402, 2009 [86]. Copyright © (2009) by the American Physical
Society.
Advances in Physics 483
an additional scattering path associated with the charge carriers in M-SWNTs. Such a scattering
path might come from the electron–electron interaction, but this theory is not yet well described.
MR,LO
ep ≡ e(k), ωLO |Hep |e(k) = gu cos R (k),
(69)
MR,TO
ep ≡ e(k), ωTO |Hep |e(k) = −gu sin R (k),
respectively, where g is the el–ph coupling constant, u the phonon amplitude and R (k) is defined
by an angle between the k vector from the K point of the 2D BZ to the van Hove singular point, kii ,
and the circumferential direction vector, K1 , [32,136,177] as shown in Figure 35(d). The values of
g are consistent with the work by Basko et al. [151,219]. Since R (k) is zero for all zigzag SWNTs
(k K1 ), we obtain MR,TO
ep = 0, while MR,LO
ep has a maximum value [196]. The meaning of R
vs. θ for SWNTs with the same family number p is shown in Figure 35(e). For the TO phonon
mode, the magnitude of the matrix element MRep for SWNTs with a similar θ value increases with
decreasing dt because of the diameter dependence in the circumferential direction [86] as shown
in Figure 35(b). The angle φ between the SWNT- axis and the phonon eigenvector for the LO
and TO phonons [280] is essential for determining the value of the el–ph matrix element. In fact,
when we consider φ, then Equation (69) is modified and becomes
Figure 35(g) and (h) show that the calculated angle φ here changes smoothly as a function of θ
[86]. The sum φLO + φTO for a general chiral angle θ always becomes π/2, because of symmetry.
The angle φ vs. θ for the LO and TO phonons can be fitted by the chiral angle dependence
(A + Bθ + Cθ 2 ) sin(6θ ), where A, B and C are fitting parameters and θ is the chiral angle in units
of degrees (◦ ). Values obtained for A, B and C for φLO , are A = 26.9, B = −76.3 and C = 84.5,
respectively, and for φTO , the corresponding values are A = −26.7, B = 75.4, and C = −83.2.
The units for the fitting parameters are degree (◦ ). This φ dependence of φLO and φTO should be
taken into account when carrying out Raman spectral calculations.
484 R. Saito et al.
in which the factor 2 comes from spin degeneracy, and Ee (k) [Eh (k)] denotes the electron (hole)
energy as a function of wave vector k, while eh(k)|Hep |ω(0) represents the el–ph matrix element
for creating an e–h pair with wave number k by the el–ph interaction Hep and f (E) is the Fermi
distribution function. The G-band spectral width is given by the decay width in Equation (71),
which is calculated self-consistently by evaluating = −Im(ω(2) ) [196,205]. The electron–
phonon interaction is used, too, for defining the ω(0) and thus we should be careful about not
double counting the constituents of this interaction [198].
Figure 36(a) shows the calculated Raman spectra for the G-band of M-SWNTs with family
number p = 30 and EF =0. The EL and γ values (see Figure 36(c)) are taken from E11 M
for each
(n, m) SWNT. The G peak intensity is larger than that of the G peak, because the G (G+ ) peak
− + −
corresponds to the LO (TO) phonon due to the LO phonon softening, in which MR,LO ep > MR,TO
ep
for any θ value, as shown in Figure 36(b). The relative intensities of the two peaks, G and G− ,
+
are affected by the Raman spectral width which relates to the phonon lifetime, . For the (10,10)
armchair SWNT, the G+ (TO) peak width is significantly smaller than those for the G− (LO) peak
Figure 36. (a) The calculated G-band spectra of M-SWNTs with the same family number p = 30 and EF =0.
(b) el–ph matrix elements vs. θ for the LO and TO phonons and for two different 2n + m family numbers.
Open-circles indicate the Mep values for the family number p = 30. (c) γ vs. θ for members of p = 30.
Reprinted with permission from J.S. Park et al., Physical Review B 80, p. 81402, 2009 [86]. Copyright ©
(2009) by the American Physical Society.
Advances in Physics 485
Figure 37. The calculated G-band spectra for three M-SWNTs with different chiral angles taken by changing
the Fermi energy from EF = −0.2 eV to 0.2 eV. (a) (10,10). (b) (11,8) and (c) (15,0). Reprinted with permis-
sion from J.S. Park et al., Physical Review B 80, p. 81402, 2009 [86]. Copyright © (2009) by the American
Physical Society.
and of the G+ peaks for the other chiral tubes. Therefore, the G+ peak intensity of the (10,10) tube
becomes large compared with the other chiral SWNTs, even though the MR,TO ep for the armchair
tube has a smaller value than that for the other chiral tubes. Since the Raman peak intensity is large
for large Mep and small values, the chiral angle dependence of these values gives an irregular
behavior to the G+ /G− spectra as a function of (n, m), as seen in Figure 36.
For a zigzag SWNT ((15, 0)), only the G+ peak appears, because MR,TO ep vanishes for zigzag
nanotubes as seen in Equation (69). The other chiral tubes in this p = 2n + m > 30 family, (11,8),
(12,6), (13,4) and (14,2), show various intermediate intensity ratios. In Figure 36(c), we show that
γ decreases monotonically with increasing θ. Because of the small difference between the γ and
the el–ph coupling for the LO phonon as compared to that for the TO phonon as a function of θ,
the G− peak intensity does not show a large change for the different chiral SWNTs. These results
show that the G-band intensity for both the G+ and G− components depends on θ, but the Raman
intensity is more sensitive to the EF position, especially for M-SWNTs.
This effect is shown more clearly by varying the Fermi Level, as shown in Figure 37(a), where
the calculated G-band spectra is plotted vs. EF at 300 K for a (10,10) armchair SWNT. Here
neither are the changes in the C–C bond nor the changes in the Eii transition energy by doping
with electrons or holes considered [86].
In Figure 37, the positive (negative) Fermi energy +EF (−EF ) corresponds to electron (hole)
doping. When EF is changed from EF = 0, the G− peak shows a frequency shift and a sharpening
of the spectral width, while the G+ peak does not show any change in intensity or width. The el–ph
interaction for the photo-excited electron does not couple to the TO phonon for armchair SWNTs
[196]. For the chiral M-SWNT (11,8) as shown in Figure 37(b), both the LO and TO phonons
couple to the intermediate e–h pair state, which is excited by a lower energy phonon. The TO
phonon becomes harder for EF = 0 eV, since the intermediate state of an e–h pair for E < ωTO
contributes to a TO phonon hardening [196]. In the case of the (15,0) SWNT, the G+ peak always
vanishes because of a vanishing MR,TO ep (See Figure 36(b)).
The matrix element Mep for the KA effect in Equation (71) is given by [196]
KA
MKA,LO
ep ≡ eh(k)|Hep |ωLO = igu sin KA (k),
(72)
MKA,TO
ep ≡ eh(k)|Hep |ωTO = −igu cos KA (k),
486 R. Saito et al.
Figure 38. (a,c,e) Experimental G-band Raman spectra which are given by the electro-chemical doping effect.
(a) Vg = 1.5 to −1.5 V. (c) Vg = 1.9 to −1.3 V. (e) Vg = 1.3 to −1.3 V with the traces taken at uniform changes
in Vg . (b,d,f) Calculated G-band Raman spectra taken by changing the Fermi energy EF in equal steps (b) 0.45
to −0.45 eV, (d) 0.60 to −0.42 eV and (f) 0.39 to −0.39 eV. The tube chiralities are: (a,b) (11,11), (c,d) (24,4)
and (e,f) (12,0). Reprinted with permission from J.S. Park et al., Physical Review B 80, p. 81402, 2009 [86].
Copyright © (2009) by the American Physical Society.
where KA (k) is defined as the angle between the k point taken on a cutting line21 for two-linear
metallic sub-bands and the nanotube circumferential direction of a unit vector, K1 . For the armchair
nanotube, the cutting line for the two-linear metallic bands lies on the nanotube axis direction unit
vector, and then KA is π/2 (−π/2), which gives a vanishing MKA,TO ep . For a chiral nanotube,
KA is not zero, since the cutting line for the two-linear metallic bands deviates from the K point
due to the curvature effect, and then the KA effect appears in both the LO and TO modes. For the
zigzag M-SWNT (15,0), only the G− peak that is related to the LO phonon appears, since the el–ph
matrix element for the Raman scattering process for iTO phonon has a zero value for a zigzag
tube, as shown in Figure 36(b). Thus, only an LO phonon softening is measured experimentally,
even though a TO phonon hardening was expected theoretically.
The calculated G-band Raman spectra vs. EF can be directly compared with the experimental G-
band Raman spectra which are obtained for electro-chemically doped individual SWNTs, as seen
in Figure 38 [86]. Here, we assume EF =0.3Vg according to Sasaki et al. [196]. The experimental
Raman spectra are shown in Figure 38(a,c,e), and the corresponding calculated Raman spectra
are shown in Figure 38(b,d,f). In Figure 38(a), the experimental Raman spectra show only a LO
phonon softening, and a TO phonon frequency shift does not occur. As mentioned above, for the
armchair SWNT, the TO phonon frequency shift does not appear and only LO phonon softening
appears. Therefore, we can predict that Figure 38(a) shows an armchair-type behavior by changing
the gate voltage. The RBM peak for these experimental Raman spectra appears at 161 cm−1 with
EL = 1.72 eV. Then we can select possible (n, m) values for a tube by using a simple tight-binding
(STB) model with γ0 = 2.9 eV for simplicity and by using the relation between the RBM frequency
and diameter, ωRBM (cm−1 ) = 248/dt (nm)22 , the possible for identifying the possible (n, m) values
for SWNTs we obtain these (n, m) values as (19,1), (18,3), (14,8) and (11,11). If our prediction
is correct, Figure 38(a) can be assigned as an (11,11) armchair SWNT. Figure 38(c) and (e) are
assigned as chiral (24,4) and zigzag (12,0) SWNTs, respectively, from the possible (n, m) values,
Advances in Physics 487
{(21, 6), (22, 4), (23, 2)} and {(10, 4), (11, 2), (12, 0)}. For the chiral M-SWNTs, not only is there
a LO phonon softening, but there is also a TO phonon hardening that appears in the calculation
of the G-band Raman spectra vs. EF . However, in Figure 38, the TO peak is too small to see on
the intensity scale of the figure. Figure 38(e) shows that the zigzag SWNT has only a G− peak
and thus only the LO phonon softening appears by changing EF , experimentally. Brown et al.
[128] and others [282,283] pointed out that asymmetric line shapes appear in the G− band Raman
spectrum for metallic tubes, which is related to the Fano resonance (Breit–Wigner–Fano, BWF line
shape) lines. Recently, Farhat et al. showed that this asymmetry is sensitive to the relative position
of the scattered light energy relative to Eii(M) , suggesting that the electron–electron interaction is
important for understanding BWF lineshapes [206].
Figure 39. Raman spectrum of single-layer graphene in comparison to graphite measured with a
Elaser = 2.41 eV (514 nm) laser. The two most intense features are named the G and G -bands. The Raman
spectrum of pristine mono-layer graphene is unique among sp2 carbons, i.e., the second-order G feature is
very intense when compared to the first-order G-band feature (see discussion in Section 4.1) [112,250].
488 R. Saito et al.
renormalization of the electronic and phonon energies, including a sensitive dependence of the
electronic structure on electron or hole doping [195]. Raman imaging can be used to define the
number of layers in different locations of a given graphene flake by measuring the dependence
of the Raman spectra (e.g., for the G-band intensity) on the number of scattering graphene layers
[294]. It is true that such information has to be analyzed with care since doping and other phys-
ical phenomena perturb the graphene Raman spectra. The effect of environmental interactions
on few-layer graphene samples have also been studied using Raman spectroscopy, including the
epitaxial growth of graphene on a substrate [295]. In this brief survey a number of important topics
are reviewed, including the spectra of mono-layer graphene, the layer number dependence in few
layer graphene, disorder-related phenomena, edge phonon phenomena, polarization effects and
the effects of doping.
Figure 40. The differences in the G Raman band for (a) 1-LG, (b) 2-LG, (c) 3-LG, (d) 4-LG, (e) HOPG
and (f) turbostratic graphite. All spectra are measured with Elaser = 2.41 eV. The original work was done by
Ferrari et al. [112] and is summarized in the review article of Malard et al. [5].
behavior discussed in connection with Figures 24 and 27. Then we turn to a discussion of the elec-
tronic properties of bi-layer graphene with AB Bernal layer stacking (as also occurs in graphite),
since the band structure change from mono-layer to bi-layer graphene is the most striking and both
mono-layer and bi-layer graphene have been probed by Raman scattering [284]. The change in the
electronic structure of graphene due to layer stacking can be probed in some detail by the DR Raman
features (see Section 3.1.3), and most sensitively by the detailed lineshape of the G -band [112].
Bi-layer graphene has a 4-peak G -band spectrum (Figure 40(b)) while mono-layer has a 1-peak
G -band (Figure 40(a)), and this fact is explained by the special electronic structure of bi-layer
graphene, which consists of two conduction and two valence bands [112], as discussed below.
Figure 41(a) shows the dispersion (peak frequency as a function of Elaser ) of each one of the four
peaks in Figure 40(b), which comprises the G -band for bi-layer graphene with AB stacking. The
double Raman resonance processes for bi-layer graphene are shown in Figure 41(b)–(e), where the
diagrams show the possible DR Raman processes that give rise to the four G peaks in Figure 41(a).
The processes are labeled by Pij (i, j = 1, 2) [284], where the states with energy Ei in the valence
band and Ej in the conduction band are connected in the photon absorption process using laser
energy Elaser . The highest frequency G peak for a given Elaser energy is associated with the P11
process, since the P11 process has the largest wave vector (q11 ) and the iTO phonon along the
KM direction in the BZ increases its frequency with increasing wave vector q (see Figure 40(b)
and Figure 41). The lowest frequency G -band peak is associated with the process P22 , which
gives rise to the smallest phonon wave vector q22 . Processes P12 and P21 shown in Figure 41(b)
give rise to the two intermediate frequency peaks of the G -band [112,284,297]. This DR Raman
490 R. Saito et al.
model has been used to relate the electronic and phonon dispersion of bi-layer graphene with the
experimental dependence of ωG on Elaser [284].
Tri-layer graphene has 15 possible DR processes [1,4], but the frequency spacings between
these peaks are not large enough to allow identification of each of the 15 scattering events
(Figure 40(c)). Increasing the number of layers increases the number of possibilities for the
G -band DR scattering processes, and an in depth analysis would get more and more complicated
for N-layer graphene (N > 3). However, experimentally the G -band spectra at a typical Elaser
energy (e.g., 2.41 eV) actually gets simpler in appearance when the number of layers increases
(see, for example Figure 40(d) for 4-LG). The spectra of increasing N converge to the two-peak
structure observed in HOPG, where N → ∞, as shown in Figure 40(e). The two-peak structure of
HOPG (Figure 40(e)) is the result of a 3D electron and phonon dispersion, as discussed in [298],
which can be seen as a convolution of an infinite number of allowed DR processes along the third
dimension of N → ∞ graphite.
Figure 41. (a) Plot of the frequency of the four Raman G -band peaks vs. Elaser observed in
bi-layer graphene. These four peaks arise from the four processes shown in (b)–(e) which com-
prise the G -band scattering processes that are expected for the phonon frequencies in bi-layer
(2-LG) graphene plotted in (a) as a function of laser energy Elaser . Reprinted with permission from
L.M. Malard et al., Physical Review B 76, p. 201401, 2007 [284]. Copyright © (2007) by the American
Physical Society.
Advances in Physics 491
and 3D graphite phases in more complicated carbon-based materials, such as pitch-based graphitic
foams [101].
The differences between stacked and non-stacked graphene layers became even more clear
when the G-band Raman spectra of AB stacked and misoriented folded bi-layer graphene were
compared [4,295,304]. While the AB stacked bi-layer graphene shows a four-peak structure, as
illustrated in Figure 40(b), the G-band spectra of misoriented bi-layer graphene shows a one-peak
profile, with an upshift of ∼14 cm−1 . This result is generally consistent with the observations
for turbostratic graphite shown in Figure 40 and it was explained as due to changes in the Fermi
velocity of graphene due to interlayer interactions in AB-stacked samples [295]. These aspects
also explain why a broadened single G peak is observed for regions of a sample that contains
domains of mono-layer or bi-layer graphene. For example, CVD-grown graphene often shows
such domains of mono-layer and bi-layer graphene and furthermore the stacking of the layers is
often not AB Bernal stacking [85,305,306]. Using a similar path of reasoning as was followed
to understand the difference in the G -band lineshape between mono-layer graphene and bi-layer
graphene with AB stacking, it is easy to understand that the G lineshape will be different for
ABC and ABA trilayer graphene stacking. Recently, Liu et al. [307] showed that the G -band can
indeed be used to distinguish between ABC versus ABA stacking in trilayer graphene samples.
The G -band for ABC stacked samples is generally broader than that for ABA, and by mapping
the G width, Liu et al. showed that these two types of stacking order coexist in trilayer graphene
samples. By mapping several samples, they showed that about 15% of the samples generated by
the mechanical exfoliation of HOPG are ABC-stacking-like, and this value for the mixed stacking
order is in very good agreement with X-ray studies on HOPG [9].
4.3. D-band and G-band intensity ratio and other disorder effects
Graphene provides an ideal structure to study the effect of disorder on a Raman spectrum, because
in a mono-layer 2D structure one does not have to worry about cascade effects and the penetration
depth of the light [3,65,292]. Here we discuss the effect of disorder caused by low energy Ar+
bombardment [3].
lattice. Introducing disorder breaks down the momentum conservation requirement, and phonons
at interior k points of the BZ can contribute to the Raman scattering process. Similar to the case
of the G -band, those scattering processes due to point defects which fulfill the DR process are
privileged in the disorder-induced Raman scattering process discussed in this section.
Finally, if the periodic structure of graphene is largely disordered, for example from a high
defect density caused by applying a large ion dose bombardment, such as 1015 Ar+ /cm2 , the
Raman spectrum evolves into a phonon DOS-like profile, where most of the higher-energy optical
phonon branch would be contributing to the spectra, rather than solely the special phonons fulfilling
the DR process (see Figure 42(c)) [3]. Of course in a fully disordered material, not only are all
phonons activated, but also changes in the structure, bonding and strain fields change the vibrational
frequencies and lineshapes.
Figure 42. The Raman spectrum of (a) crystalline graphene, (b) defective graphene, (c) and fully disordered
single-layer graphene. These spectra were obtained with Elaser = 2.41 eV and the graphenes are deposited
on a SiO2 substrate using the mechanical exfoliation method (scotch-tape). Reprinted with permission from
A. Jorio et al., Journal of Physics: Condensed Matter 22, p. 334204, 2010 [309]. Copyright © (2010) by the
Institute of Physics.
Advances in Physics 493
Figure 43. Evolution of the first-order Raman spectra using a λ = 514 nm laser (Elaser = 2.41 eV) to investi-
gate a graphene mono-layer sample deposited on an SiO2 substrate, and subjected to Ar+ ion bombardment.
The Ar+ ion doses from the bottom trace to the top trace are: zero (pristine), 1011 , 1012 , 1013 and 1014
Ar+ /cm2 for 90 eV ions. The spectra in this figure are also displaced vertically for clarity. Reprinted From
Carbon 48(5), M.M. Lucchese et al., pp. 1592–1597 [3]. Copyright © (2010) Elsevier.
494 R. Saito et al.
Figure 44. The ID /IG data points from three different mono-layer graphene samples as a function of the aver-
age distance LD between defects, induced by the Ar+ ion bombardment procedure described in Section 4.3.2.
The solid line is a modeling of the experimental data with Equation (73). The inset shows a plot of ID /IG vs.
LD on a log scale for two samples: (i) a ∼50-layer graphene sample; (ii) a 2 mm-thick HOPG sample, whose
measured values are here scaled by (ID /IG ) ×3.5. Reprinted From Carbon 48(5), M.M. Lucchese et al., pp.
1592–1597 [3]. Copyright © (2010) Elsevier.
Figure 45. (a) Definition of the “activated” A-region (darkest gray) and “structurally disordered” S-region
(dark gray). The radii rS and rA are measured from the impact point which is chosen randomly in this
simulation. (b–e) show 55 nm×55 nm portions of the graphene simulation cell, with snapshots of the structural
evolution of the graphene sheet for different defect concentrations: (b) 1011 Ar+ /cm2 ; (c) 1012 Ar+ /cm2 ;
(d) 1013 Ar+ /cm2 and (e) 1014 Ar+ /cm2 , as in Figure 43. Reprinted From Carbon 48(5), M.M. Lucchese et
al., pp. 1592–1597 [3]. Copyright © (2010) Elsevier.
most strongly to the D-band, while the S-regions will make less contribution to the D-band due
to the break-down of the lattice structure itself. These two different scattering cross sections for
the disorder-induced processes will give rise to the non-monotonic behavior observed in the LD
dependence of the ID /IG ratio, as shown in Figure 44.
The structurally disordered (S) region and the activated (A) region are shown in Figure 45(a) by
light and dark gray regions, respectively. The evolution of the S and A regions for a graphene sheet
under ion bombardment was simulated by randomly choosing a sequence of impact positions on
a graphene sheet. As the number of impacts increase, the activated A-region increases, leading to
a decrease in LD and an increase of the D-band intensity ID . When the graphene is fully covered
with A-regions, an increase in ion bombardment fluence causes the structurally disordered S-
regions to take over from the A-regions, thus leading to a decrease of the D-band intensity ID
(see Figure 45(b–e)). This model is the basis for the evolution of ID /IG based on Equation (73)
which, with the parameters CA = 4.56, CS = 0.86, rA = 3 nm and rS = 1 nm, give the line curve
Advances in Physics 495
in Figure 44 that fully describes the experimental evolution of ID /IG , shown by the black bullets
in Figure 44 [3].
For low defect concentrations (large LD values), ID /IG = (102 ± 2)/LD2 , which means the total
area contributing to scattering is proportional to the number of defects. This regime is valid for
LD > 2rA , while below this limit for LD , the activated regions start to overlap (see Figure 45(e)),
thus changing the simple ID /IG ∝ LD−2 dependence. The D-band intensity then reaches a maximum
and a further increase in the defect concentration decreases the D-band intensity because the
graphene sheet starts to be dominated by the structurally disordered areas (S-region).
The rS = 1 nm value is in agreement with the average size of the disordered structures seen
in the STM images [3,130]. This is not a universal parameter, but is a parameter that is actually
specific to the ion bombardment process. The φ = rA − rS = 2 nm value represents the Raman
relaxation length for the defect-induced resonant Raman scattering in graphene. This value is valid
for the laser excitation energy 2.41 eV and room temperature, and may change with changing Elaser
and temperature. Be aware that this is the relaxation length for the excited electrons, which should
not be confused with the relaxation length for the phonons. The value CA = 4.56 is in rough
agreement with the ratio between the electron–phonon coupling for the iTO phonons evaluated
between the and the K points in the BZ [218–220], which is consistent with the expectation that
the CA parameter should be related to the electron–phonon matrix elements. The CS parameter is
related to the size of the highly disordered area, and there is no theoretical work yet available on
this matter.
It is important to have an equation relating ID /IG to LD that can be used by researchers looking
for a Raman characterization of the defect density present in a specific graphene sample. The
entire regime (0 → LD → ∞) can be fitted using [3]:
ID r 2 − rS2 −π rS2 −π(rA2 − rS2 ) −π rS2
= CA 2A exp − exp + C S 1 − , (74)
IG rA − 2rS2 LD2 LD2 LD2
which comes from solving rate equations for the bombardment process. Fitting the data in Figure 44
with Equation (74) gives CA = (4.2 ± 0.1), CS = (0.87 ± 0.05), rA = (3.00 ± 0.03) nm and rS =
(1.00 ± 0.04) nm. This equation represent the results very well, since the fitting obtained with
Equation (74) is also in very good agreement with experiment and the fitting parameters are fully
consistent with the parameters obtained by computational modeling using Equation (73) [3].
4.3.4. The Local Activation Model and the Raman Integrated Areas
The dependence of the intensity ratio ID /IG on LD was found to accurately follow an analytical
formula (Equation (74)), as described above, and this result is useful for practical applications and
for inter-laboratory comparisons. However, the physics behind this effect has to take into account
that both ID and IG vary when LD is changed. As discussed in Section 1.4.5, the evolution of the
Raman profile can be discussed as related to the peak intensity or to the integrated peak area.
In this section, we choose to use the same model as was used to derive Equation (74) when we
analyze the evolution of the intensity and integrated area of the many Raman peaks that vary with
increasing structural disorder, by normalizing each of them to the G-band integrated area (see
Figure 46). As shown in the inset to Figure 46 (top-right panel), the integrated area of the G-band
does not show any simple evolution with disorder [130].
The lower-left panels of Figure 46 show that this analytical expression fits the quantities
AD /AG and AD /AG nearly perfectly, where A refers to the integrated area. For the D-band, the
fitting parameters are rS = 2.6 nm, rA = 4.1 nm, CS = 2.4 and CA = 3.6, whereas for the D -
band the fitting parameters are rS = 2.6 nm, rA = 3.8 nm, CS = 0.28 and CA = 0.19. Note that
we obtain close to the same value of rS for both the D and D modes, indicating that indeed rS
496 R. Saito et al.
Figure 46. Normalized intensities (upper panel) and areas (lower panel) of the Raman D-, D -, G-
and G -bands as a function of LD . All quantities are normalized by the area of the G-band (see
the as-measured AG in the inset to the upper-right panel). The solid lines in the lower panel are
theoretical results based on the model described in Section 4.3.4. Reprinted figure with permission
from E.H. Martins Ferreira et al., Physical Review B 82, p. 125429, 2010 [130]. Copyright © (2010) by the
American Physical Society.
is a geometrical, structure-related length. Also, we find 1.5 and 1.3 nm for the spatial extent of
the Raman processes rA − rS , which is of the same order of magnitude as the rough estimates
vF /ωD = 4.3 nm and vF /ωD = 3.6 nm. We remind the reader that the distance rA − rS is a rough
measure of the length traveled over the lifetime of the e–h pair, vF /ωX , where vF is the graphene
Fermi velocity of the electron and hole carriers and ωX is the frequency of any X phonon mode
[130]. More interestingly, the ratio between rA − rS for the D- and D -bands matches very closely
to the ratio of the inverse frequencies ωD /ωD ≈ 1.2.
Similar ideas can be applied to a discussion of the AG /AG ratio, but in this case, since the G -
band is already active for pristine graphene, the intensity ratio is only affected by the disruption
of the hexagonal network, leading to a decrease in the AG /AG ratio as a function of increasing
Advances in Physics 497
where AG /AG (∞) is the area ratio for pristine graphene while AG /AG (LD ) is the area ratio for
an actual sample characterized by its LD value. The fitting of the experimental data, shown in
the lower-right panel of Figure 46, gives in this case rS = 2.5 nm, which is also similar to the
structural damage length obtained for the D- and D -band spectra. This result is in accordance
with the typical defect-size estimates found independently from the STM analysis [3,130]. In
Section 4.3.5, we describe what happens to the frequency and linewidth of the Raman peaks as a
result of ion implantation-induced structural damage.
4.3.5. Modeling disorder effects in the Raman linewidths and frequency shifts: the spatial
correlation model for defects
Disorder introduced by a random distribution of defects causes a broadening and a shifting of
the Raman mode frequencies and increases in the asymmetry of both the Raman-allowed and
the newly disorder-activated Raman bands discussed in Section 4.3.4. Here, we use the so-called
“spatial-correlation model” introduced by Capaz and Moutinho in [130] to describe these effects
in graphene. Other work on this topic that should also be referred to is in Refs. [65,292],
As described in Section 1.4.6, a random distribution of point defects will scatter phonons and it
will also add a contribution to the FWHM by an el–ph coupling of phonons with wave vectors q0
and q0 + δq. In the limit of low levels of disorder, the Raman intensity for the disordered graphene
I(ω) can be calculated by Equation (7). With this model, we can calculate the full lineshape of I(ω)
and from that we can extract the disorder-induced peak shifts ωq0 (Figure 47, lower panel) and
the increases in the FWHM q0 (Figure 48, lower panel). Since we use experimentally available
dispersion relations ω(q), the only fitting elements in this model are: (1) the relationship between
the coherence length L and the typical inter-defect distance LD , and (2) the weighting function
W (q) in Equation (7).
We now describe in more detail the application of the above model to the different Raman bands
considered in graphene, including the G-band, the D -band, the D-band and the G -band [130].
A. G-band – The G-band in perfect graphene is associated with phonons at the -point, i.e., q0 =
0 phonons. We consider that disorder mixes equally the -point phonons with nearby phonons in
both the LO and iTO phonon branches. We find that the best agreement with experiment is obtained
by using a constant weighting function (which is equivalent to not use a weighting function at all).
For the LO and TO phonon dispersions, we take
where ωn (q) is in cm−1 (n = LO or iTO) and ωG = 1580 cm−1 is the experimental G-band frequency
for pristine graphene used in this work. Here q is measured from the -point in units of Å−1 . These
dispersions are taken from the work of Maultzsch et al. [208] by interpolating the frequencies at
high-symmetry points and by averaging the dispersions between the –K and –M directions.
Also, since the main contribution to the integral in Equation (7) will come from q vectors near the
point, the BZ can be safely approximated by a circular disk and the integral will be considered
explicitly in the radial coordinate only. Taking all these considerations into account, Equation (7)
498 R. Saito et al.
Figure 47. The upper panel shows peak frequencies of the D, G, D and G -bands as a function of LD denoting
a typical distance between defects. The inset compares the frequency of the D-band and the G -band divided
by two, showing that we always have ωG /2 < ωD , in agreement wtih Ref. [216]. The lower panel shows
frequency shifts with respect to the zero-disorder limit. Dots are experimental points and solid lines are theo-
retical results based on the model described in the text. Experimental error bars are 2 cm−1 . Reprinted figure
with permission from E.H. Martins Ferreira et al., Physical Review B 82, p. 125429, 2010 [130]. Copyright
© (2010) by the American Physical Society.
becomes [130]
exp[−q2 L 2 /4]
IG (ω) ∝ 2π qdq (77)
n
[ω − ωn (q)]2 + [0 /2]2
in which the sum is over the two (LO and iTO) phonon branches.
B. D -band – The D -band arises from intra-valley phonons with a linear wavevector intensity
dependence with respect to the laser energy. Since the D -band has been assigned to LO phonons,
only this branch is considered in calculations of the D -band intensity using Equation (7). We
Advances in Physics 499
Figure 48. (a) FWHM intensity of the D-, G-, D - and G -bands as a function of LD , denoting the typical
distance between defects. (b) Disorder contribution to the peak widths, , for the D, G, D and G -bands.
Points denote are experiments and solid lines are theoretical results based on the model described in the text.
Reprinted figure with permission from E.H. Martins Ferreira et al., Physical Review B 82, p. 125429, 2010
[130]. Copyright © (2010) by the American Physical Society.
average over all possible directions θ of the wavevector q0 and, similarly to the case of the G-
band, there is no need to introduce a q-dependent weighing function W (q). Then, the D -band
intensity becomes [130]:
exp[−(q − q0 )2 L 2 /4]
ID (ω) ∝ qdq dθ . (78)
[ω − ωLO (q)]2 + [0 /2]2
For the laser energy of 2.41 eV, the value for |q0 | in Equation (78) is found to be |q0 | = 0.42 Å−1
measured from the point.
C. D-band – The D-band arises from inter-valley phonons which also show a linear wavevector
dependence with respect to the laser energy. In fact, for the laser energy of 2.41 eV, we also find
|q0 | = 0.42 Å−1 for the D-band, but now q0 is measured from the K point. Since the D-band
has been assigned to iTO phonons along the K–M direction in the BZ, we choose q0 along this
500 R. Saito et al.
direction and the weighting function W (q) is also restricted to be non-zero only along the same
direction. Mathematically, W (q) = δ(θ − θK−M )f (q), where θK−M indicates the K–M direction
and f (q) = 1 + a(q0 − q) is a function that linearizes the radial dependence of the electron–
phonon coupling along the K–M direction near q0 . With these conditions, the D-band intensity
becomes [130]
f (q) exp[−(q − q0 )2 L 2 /4]
ID (ω) ∝ dq . (79)
[ω − ωiTO (q)]2 + [0 /2]2
For the iTO phonon dispersion along the K–M direction, we use [244]:
where ωiTO is in cm−1 and q is measured from the K point in units of Å−1 .
D. G -band – The G -band is related to a DR process associated with the same inter-valley
phonons as the D-band. For this reason, the expression for the intensity becomes more complicated
and it involves a double integral over the forward (q) and backward (q ) phonon wavevectors. Using
the same considerations for the el–ph matrix elements, which essentially select phonons in the
K–M direction, we have
!
"
f (q)f (q ) exp (−[(q − q0 ) + (q − q0 ) ]L )/4
2 2 2
IG (ω) ∝ dq dq (81)
[ω − ωiTO (q) − ωiTO (q )]2 + [0 /2]2
where f (q) is the same linear function as in the D-band case and q0 is also the same. We also
impose the condition that the same relation between L (the disorder-induced phonon coherence
length) and LD (the average distance between defects) must be valid for the D and G -bands.
In Figures 47 and 48, we see the results for the frequency and linewidth as a function of
the typical distance LD between defects for the data fitting of the frequency shifts and widths,
respectively, as described above. Note that the general agreement is good, especially for large
values of LD . Indeed, this spatial correlation model, because of its perturbation character, is not
expected to be valid in the highly disordered regime. In Figures 47 and 48, the best relationships
between L and LD in each case are shown (as obtained by the fits between the model and the
experimental data). It is physically reasonable to see that L and LD are similar to each other. This
condition was not imposed, but it comes automatically from the fitting procedure. This means that
the disordered-induced phonon coherence length L is of the same order of magnitude as the typical
inter-defect distance LD , which is physically reasonable. There is no reason to expect that the same
relation between L and LD should be found for the different phonon modes, since different modes
should have different defect scattering cross-sections. From the results shown here, it seems that
the D modes are the most affected by point defect disorder, showing a smaller coherence length
than the other modes for the same amount of disorder. Finally, the model allows us to explain the
greater increase in the FWHM for each of the modes near the K point relative to the modes near
the point as being simply a consequence of the larger magnitude of the phonon dispersions near
the K point.
Figure 49. Evolution of the G -band (at 2670 cm−1 ) and other second-order peaks, the (D + G) at 2930 cm−1 ,
the (D + G) at 3190 cm−1 , and the G (2D in the figure) at 3220 cm−1 with increasing ion doses. The
intensities of the two lower graphs are multiplied by a factor of 10 for the sake of readability [130]. Here the
notation 2D is used instead of G , as has also been used in the literature by other authors. Reprinted figure
with permission from E.H. Martins Ferreira et al., Physical Review B 82, p. 125429, 2010 [130]. Copyright
© (2010) by the American Physical Society.
1014 Ar + /cm2 the results show a frequency downshift for all DR features, in agreement with the
results of Section 4.3.5.
experimentally, thereby distinguishing the so-called zigzag edge from the armchair or random
atomic edge structures [170]. The armchair/random vs. zigzag edge structure can be identified
spectroscopically by the presence vs. absence of the D-band, and this effect results from the
momentum requirements of the DR model, as discussed below.
The defect associated with a step edge has a 1D character, which means that it is able to transfer
momentum solely in the direction perpendicular to the edge. In this sense, the wave vectors of
the defects associated with zigzag and armchair edges are represented in Figure 50(a) by d a (a
for armchair) and d z (z for zigzag) edges. When we translate these vectors into reciprocal space,
we see that different selection rules apply for the electron scattering by phonons for each of these
edge types. This is illustrated in Figure 50(b), where the first BZ of 2D graphite (graphene) is
shown, oriented in accordance with the real space directions shown in Figure 50(a).
Light-induced e–h pairs will be created on an equi-energy circle around points K and K (here
neglecting the trigonal warping effect for simplicity), which has a radius that is defined by Elaser , as
shown in Figure 50(b). Note that for inter-valley electron-defect scattering, which connects K to
K points, only the d a vector for armchair edges can connect points belonging to circles centered at
two inequivalent K and K points. In contrast the zigzag d z vector to connect inequivalent points,
which means that inter-valley scattering is not allowed for zigzag edges. This therefore means that
the inter-valley DR process, which is the process responsible for the observation of the D-band in
graphitic materials, is not allowed for a perfect zigzag edge [170]. The D-band phonon connects
two inequivalent K and K points, and along the zigzag edge there will be no defect able to connect
those points to achieve momentum conservation in the final process.
On the other hand, intra-valley electron-defect scattering can occur for both zigzag and armchair
edges (see Figure 50(b)). Therefore, intra-valley scattering processes induced by phonons can
achieve final momentum conservation using both d a and d
z vectors. Another well-known defect-
induced band is the so-called the D -band, which appears at around 1620 cm−1 , and it is related
to intra-valley el–ph processes. For this reason, the D -band observation should be independent of
the zigzag vs. armchair structure of the edges, in agreement with experimental observation.
Another selection rule aspect refers to the D-band intensity dependence on the polarization
direction of the light with respect to the edge orientation. The D-band intensity has a maximum
value when the light is polarized along the edge, and should give a null value when the light is
polarized perpendicular to the edge. The physics behind this selection rule is the optical absorption
(emission) anisotropy around the K(K ) point in 2D graphite, which can be represented by [262]
Wabs,ems ∝ |P 2.
× k| (82)
Here the polarization of the incident (scattered) light for the absorption (emission) process is
represented by P
, while the wave vector of the electron measured from the K point is given by k.
These selection rules were first observed for graphite edges, as reported in [170], and similar
results have been observed later in mono-layer graphene [151,311]. However, only edge-dependent
variations in the D-band intensity consistent with the selection rules have been reported. Raman-
based indications for the high crystallinity of zigzag edges have indeed been observed by Krauss
et al. [312], although the complete absence of the D-band together with the observation of the
D -band, which is expected for a zigzag edge structure, has never been reported, which might
imply that, up to now, no perfect zigzag structure has been measured by Raman spectroscopy. In
general, the polarization direction dependence for the D-band intensity, as given by Equation (82),
together with the zigzag vs. armchair dependence, can be used for an identification of the edge
orientation and structure. Raman spectroscopy is, therefore, a valuable tool for the development
of our understanding of edge structures, important for the science of graphene ribbons, and more.
The results reported here represent an effort to improve our understanding of the influence of
Advances in Physics 503
Figure 50. (a) Schematic illustration of the atomic structure of edges with the zigzag and armchair ori-
entations. The boundaries can scatter electrons with momentum transfer along d z for the zigzag edge,
and along da for the armchair edge. (b) First BZ of 2D graphite (graphene), showing defect-induced
z is too short to connect the K and K points,
inter-valley and intra-valley scattering processes. Since d
the defect-induced DR inter-valley process is forbidden at zigzag edges. Reprinted figure with permission
from L.G. Cancado et al., Physical Review B 93, p. 47403, 2004 [59]. Copyright © (2004) by the American
Physical Society.
the specific defect structure on the Raman spectra of sp2 carbon systems. Other defect-dependent
effects are expected, which may be very useful to characterize defects in nanographite-based
devices, but both theory and experiment have to be developed along these lines.
Figure 51. (a) The G-band Raman spectra from a graphene nano-ribbon (G1 ) and from the HOPG substrate
(G2 ) on which the nano-ribbon was grown. (b) The dependence of the G1 frequency on the light polarization
direction, with respect to the ribbon axis. Points are experimental results and the dashed curve is the theoretical
expectation. (c) Frequency of the G1 and G2 peaks as a function of the incident laser power density. Reprinted
figure with permission from L.G. Cancado et al., Physical Review B 93, p. 47403, 2004 [59]. Copyright ©
(2004) by the American Physical Society.
Advances in Physics 505
nanotubes and, therefore, it can be used to distinguish SWNTs from other sp2 carbon structures in
the samples. In general the intensity of the RBM is unusually strong when compared with other
non-resonant spectral features coming from other carbonaceous materials or from the substrate
on which the tubes are sitting [111]. Furthermore, the RBM frequency ωRBM depends on the tube
diameter, following the proportionality relation ωRBM ∝ 1/dt . This dependence was predicted
initially using force constant calculation models (e.g. [100]), but a rather simple and instructive
analytical derivation can be made using elasticity theory, and we present this approach in sequence.
ρ ∂ 2 x(t) 2 (1 − ν 2 )
(1 − ν 2 ) 2
+ x(t) = − p(x), (83)
Y ∂t dt Yh
where x(t) is the displacement of the nanotube in the radial direction, p(x) = (24K/s02 )x(t),
and K (in eV/Å2 ) gives the van der Waals interaction strength, s0 the equilibrium separation
between the SWNT wall and the surrounding environmental shell, Y the Young’s modulus
(69.74 × 1011 g/cm·s2 ), ρ the mass density per unit volume (2.31 gm/cm3 ), ν = 0.5849 is the
Poisson’s ratio and h represents the thickness of the environmental shell [315]. If there are no
environmental effects, the term p(x) vanishes and Equation (83) will become the fundamental
0
frequency ωRBM for a pristine SWNT in units of cm−1 ,
# 1/2 $
1 Y 1
ωRBM =
0
. (84)
πc ρ(1 − ν ) 2 dt
The term inside the curly bracket above gives the fundamental value of A = 227.0 cm−1 nm.
However, for a non-vanishing inward pressure p(x), the result is
1/2
1 6(1 − ν 2 ) K
ωRBM = 227.0 2 + . (85)
dt Yh s02
Here [6(1 − ν 2 )/Yh] = 26.3 Å2 /eV. The shift in ωRBM due to the nanotube environment is given
by ωRBM = ωRBM − ωRBM . We fit the value K/s02 in Equation (3) to the RBM frequency as
0
a function of dt . The fitted value for the environmental term for the “super-growth” sample is
sufficiently small, since K/s02 = (2.2 ± 0.1) meV/Å4 , that it can be neglected. The dt dependent
506 R. Saito et al.
Ce Sample Reference
behavior of the environmental effect in ωRBM reproduces well the experimental result for dt up to
dt = 3 nm [314]. A similar environmental effect is obtained for SWNTs surrounded by different
surfactants [107,175,188,317,319], in bundles [109,321], sitting on a SiO2 substrate [111], and
even for tubes suspended in air by posts [320]. This environmental effect is almost absent in
“super-growth” SWNTs, but the reason why this sample is special is not presently understood,
and the pristine-like ωRBM behavior is lost if this sample is dispersed in solution [322].
A simple relation can be proposed for all the ωRBM results in the literature, which are generally
upshifted from the pristine values observed for the “super-growth” samples due to the van der
Waals interaction with the environment. This simple relation is [314]
%
227
Lit.
ωRBM = 1 + Ce ∗ dt2 , (86)
dt
where Ce in Equation (86) represents the effect of the environment on ωRBM , i.e. Ce = [6(1 −
ν 2 )/Eh][K/s02 ] nm−2 . The several Ce values that are obtained by fitting the RBM results for
different commonly found samples in the literature are given in Table 2. The curvature effects
become important for dt < 1.2 nm, and in this case the environmental effect depends more critically
on the specific sample. For example, the Ce for SWNT samples sitting on a SiO2 substrate may
differ from sample to sample. The observed environmentally induced upshifts for the RBM from
small diameter tubes, either within bundles or wrapped by different surfactants (e.g., SDS (sodium
dodecyl sulfate) or single stranded DNA), range from 1 to 10 cm−1 . This environmental effect gets
richer in a double wall carbon nanotube (DWNT), as discussed in Section 5.1.2.
Finally, all the ωRBM dependence on the carbon nanotube structure discussed here addresses
the importance of the diameter dependence. The chiral angle has a weaker dependence on the
RBM frequency, but to fully discuss this topic, the KA has to be introduced [193,324]. This topic
is discussed in Section 5.4.
which specific inner and outer tubes form a given DWNT, one has to perform Raman experiments
on individual DWNTs (see Figure 52(c)).
For a well-defined experiment, the combination of electron-beam lithography, atomic force
microscopy (AFM) and Raman spectral mapping have been developed to measure the Raman
spectra from the inner and the outer tubes of an individual DWNT (see Figure 52) [330,331]. The
Raman spectra of 11 isolated DWNTs grown from annealing C60 filled SWNTs were measured
using a single laser excitation energy (Elaser = 2.10 eV [330]).Specific Elaser values were used to
select all DWNTs with (6,5) semiconducting inner tubes that were in resonance, and all with the
S@M configuration so that the RBMs from both the inner and outer tubes of individual DWNTs
could be observed. The RBM frequencies ωRBM,o for the outer tube measured for such a DWNT as
a function of ωRBM,i for the inner tube are shown in Figure 53(a). For these 11 individual isolated
DWNTs, ωRBM,o for the outer tubes varies along with ωRBM,i , thus showing that the inner and outer
tubes impose considerable stress on one another.Actually, the nominal wall-to-wall distances dt,io
between the inner (i) and outer (o) tubes of the DWNTs are less than the 0.335 nm interlayer c-axis
distance in graphite. Figure 53(b) shows dt,io values as small as 0.29 nm, with a decrease of up
to 13% in the wall to wall distance [330]. Because of the differences in the Coulomb interaction
Figure 52. (a) Raman spectra for the RBM region for two types of DWNTs, obtained with Elaser = 2.13 eV.
(b) AFM image of one individual DWNT. The inset shows the silicon substrate with gold markers showing
the location of an individual DWNT. (c) RBM Raman spectra obtained with Elaser = 2.11 eV for an isolated
individual DWNT grown from a C60 filled SWNT-bundle. (d) AFM height profile of the individual, isolated
DWNT shown in (b), with the RBM spectrum shown in (c). The vertical lines connecting (a) and (c) show
that the ωRBM of the prominent tube diameters observed in the C60 -DWNT bundles coincide with the ωRBM
of the inner and outer tubes of the isolated C60 -DWNT. F. Villalpando-Paez et al., Nanoscale 2, pp. 406–411,
2010 [330]. Adapted by permission of the Royal Society of Chemistry.
508 R. Saito et al.
expected for the four different DWNT metallicity configurations, S@M, M@S, S@S and M@M,
the detailed relation between ωRBM and 1/dt will depend on the metallicity configuration.
By adding walls to form MWNTs, the RBM signal from inner tubes with small enough diam-
eters (dt 2 nm) can be observed experimentally [332]. However, most of the usually made
MWNT samples are composed of inner tubes with diameters too large to exhibit observable RBM
features.
Figure 53. All the inner tubes for the 11 peapod-DWNTs in this figure are (6,5) semiconducting tubes. (a)
Plot of the ωRBM,i for the inner tube vs. ωRBM,o for the outer tubes which pair to form eleven different isolated
DWNTs. (b) Plot of the nominal wall-to-wall distance dt,io for each of the 11 isolated DWNTs vs. ωRBM,i
shown in (a). An increase in the ωRBM,i of the (6,5) inner tubes shown here is accompanied by a decrease
in the measured nominal wall-to-wall dt,io distance for these peapod-derived DWNTs. F. Villalpando-Paez
et al., Nanoscale 2, pp. 406–411, 2010 [330]. Adapted by permission of the Royal Society of Chemistry.
Advances in Physics 509
The Stokes RBM peak intensity I is a function of Elaser and can be evaluated from Equation (62)
[269]. The two factors in the denominator of Equation (62) describe the resonance effect with
the incident and scattered light, respectively. A ± sign before Eph applies to Stokes/anti-Stokes
processes, while γRBM is related to the inverse lifetime for the resonant scattering process [278]. The
matrix elements in the numerator are most usually considered to be independent of energy because
of the small energy range. The theory for these matrix elements is discussed in Section 3.6.1. The
lines in Figure 55 show the fits to the experimental data for the Stokes (dashed) and anti-Stokes
(solid) resonance windows, using Eph = 21.5 meV, obtained from ωRBM = 173.6 cm−1 [338]. The
asymmetric lineshape in the resonance windows in Figure 55 was obtained in [338] by considering
not a coherent & Raman process, but an incoherent scattering process, where the sum over the
internal states ( a in Equation (62)) was taken outside the square modulus. This procedure is
indeed controversial because this asymmetry could be also generated by different γr values for
the incident and scattered resonance windows or by other resonance levels lying close in energy.
Disregarding the asymmetry aspect, Eii = 1.655 ± 0.003 eV and γRBM = 8 meV is obtained
for the spectra shown in Figure 55. A shift in the Stokes (S) and anti-Stokes (aS) resonant windows
is expected due to the resonant condition for the scattered photon, Es = Eii ± Eph , with (+) for
the Stokes and (−) for the anti-Stokes processes, and this effect is shown in the upper inset to
Figure 55. For this reason, under sharp resonance conditions the IaS /IS intensity ratio depends
sensitively on Eii − Elaser , and the IaS /IS ratio can be used to determine Eii experimentally and to
determine whether the resonance is with the incident or the scattered photon [141,338].
Figure 54. (a) AFM image of a SWNT sample showing markers that were used to localize the spot
position (dashed circle) on the substrate during the Raman experiment and for further AFM charac-
terization of the SWNTs that are located within the light spot indicated by the dashed circle in (b).
(c) anti-Stokes and (d) Stokes Raman spectra from isolated SWNTs on a Si/SiO2 substrate for sev-
eral different laser excitation energies. For more details, see Ref. [338]. Adapted with permission from
A. Jorio et al., Physical Review B 63, p. 245416, 2001 [338]. Copyright © (2001) by the American Physical
Society.
510 R. Saito et al.
Figure 55. Raman intensity vs. laser excitation energy El for the ωRBM = 173.6 cm−1 peak in Figure 54, for
both anti-Stokes and Stokes processes. Circles and squares indicate two different runs on the same sample.
The line curves indicate the resonant Raman window&predicted from Equation (1), with Eii = 1.655 eV,
γr = 8 meV, but taking the sum over internal states ( m,m ) outside the square modulus. The upper inset
compares the theoretically predicted Stokes and anti-Stokes resonant windows on an energy scale in eV, and
the lower insert shows the joint density of states (JDOS) vs. Elaser for this SWNT. Adapted with permission
from A. Jorio et al., Physical Review B 63, p. 245416, 2001 [338]. Copyright © (2001) by the American
Physical Society.
Figure 56. (a) Raman spectrum (bullets) of SWNT bundles obtained with a 644 nm laser line
(Elaser = 1.925 eV). This spectrum was fitted by using 34 Lorentzians (curves under the spectra) and the
solid line is the fitting result. (b) The Kataura plot used as a guide for the fitting procedure. Adapted with
permission from P.T. Araujo et al., Physical Review Letters 98, p. 67401, 2007 [321]. Copyright © (2007)
by the American Physical Society.
Advances in Physics 511
Figure 57. Resonance windows for specific (n, m) SWNTs within a bundle. (a) Resonance profile (black
dots) in the near-infrared range for ωRBM = 192.7 cm−1 . The data for tube (14, 3) were fitted (solid line)
using Equation (87) with γ = 0.065 eV and Eii = 1.360 eV. (b) Resonance profile in the visible range is
shown for ωRBM = 192.5 cm−1 (tube (12, 6)), with γ = 0.045 eV and Eii = 1.920 eV. (c) Resonance pro-
file in the near-ultraviolet range is shown for ωRBM = 257.6 cm−1 (tube (11, 1)), with γ = 0.073 eV and
Eii = 2.890 eV.Adapted with permission from P.T. Araujo et al., Physical Review Letters 98, p. 67401, 2007
[321]. Copyright © (2007) by the American Physical Society.
and the green Lorentzians represent the RBMs from semiconducting SWNTs. The number of
Lorentzians used to fit each resonance spectrum can be defined by the Kataura plot [177,342,343]
(see Figure 56(b)), which plots the Eii for all possible (n, m) SWNTs. The ωRBM values obey
the relation ωRBM = (227/dt ) 1 + Ce /dt2 , which correctly describes environmental effects by
the proper choice of Ce , and this relation is discussed in detail in Section 5.1.1. For lack of
information, we assume that all the Lorentzian peaks in one experimental spectrum share the
same FWHM value.
The RBM peak intensity I(Elaser ) for each Lorentzian peak in the RBM spectra from an
individual SWNT can be evaluated by
2
1
I(Elaser ) ∝ ,
(87)
(E laser − Eii − iγ )(Elaser − Eii ± Eph − iγ )
which is a simplification of Equation (62). Figure 57 shows the resonance profiles for three
different (n, m) SWNTs [321,335] The resonance window widths γr for SWNTs in bundles are
usually within the 40–160 meV range, which are much broader than for isolated SWNTs (see
Figure 55) and γr also depends on (n, m) [321].
equally abundant in the growth process. Actually, chiral SWNTs are twice as populous as achiral
ones because there are right-handed and left-handed isomers present in a typical sample. The
intensity calibrated experimental RRS map in shown in Figure 58(a). The (n, m) nanotubes in the
(2n + m) = constant family have similar diameters and Eii values to one another, and the Raman
intensity within a given (2n + m) = p = constant family has a chiral angle dependence. From the
spectral map it is clear that the RBM intensity is stronger for smaller chiral angles (near zigzag
nanotubes) as compared to those with larger chiral angles. Each spectrum (S(ω,EL ), ) is the sum of
the individual contributions of all SWNTs present in the light beam, and we can write [344]
/2
S(ω,EL ) = Pop(n, m)I(n, m)EL , (88)
n,m
(ω − ωRBM )2 + (/2)2
where Pop(n, m) is the relative population of the (n, m) nanotube species, = 3 cm−1 is the
experimental average value for the FWHM intensity of the Raman spectra (Lorentzian), ωRBM is
the RBM frequency and ω is the corresponding Raman shift variable. The total integrated area
(I(n, m)EL ) for the Stokes process at a given excitation laser energy (EL ) is given by
2
M
I(n, m)EL = , (89)
(EL − Eii + iγ )(EL − Eii − Eph + iγ )
where the superscript EL denotes the laser excitation energy, Eph = ωRBM is the energy of the
RBM phonon, Eii is the energy corresponding to the ith excitonic transition, γ is the resonance
window width and M represents the matrix elements for the Raman scattering by one RBM
phonon of the (n, m) nanotube. The values for Eii and ωRBM have to be determined experimentally.
The effective matrix element square term M and the effective resonance window width γ for
each (n, m) tube were found by fitting the experimental RBM component of the RRS map with
Equation (88) using the functions:
MB MC cos(3θ ) 2 γB γC cos(3θ)
M = MA + + and γ = γA + + . (90)
dt dt2 dt dt2
Figure 58. (a) Experimental RRS map for the RBM feature. The intensity calibration was made by measuring
a standard tylenol sample. (b) Modeled map obtained by using Equation (88) in the same laser excitation
energy range as (a). Adapted with permission from P.B.C. Pesce et al., Applied Physical Letters 96, p. 51910,
2010 [344]. Copyright © (2010) by the American Institute of Physics.
Advances in Physics 513
Here Mi and γi (i = a, b, c) are fitting parameters. The best values for Mi and γi , considering
S M
the excitonic transitions E22 and the lower branch of E11 , are listed in Table 3. Here dt is given in
nm, γ in meV and M in arbitrary units.
The modeled RRS map shown in Figure 58(b) was obtained using the values thus obtained in
Equation (88). Note that the model represents the experimentally observed results in Figure 58(a)
very well.
Type MA MB MC γA γB γC
Figure 59. (a) RBM resonance Raman map for the “super-growth” (S.G.) SWNT sample [314,335,345].
(b) Kataura plot of all transition energies (EiiS.G. ) that could be experimentally obtained from the resonance
windows extracted from (a) as a function of ωRBM . (c) Kataura plot obtained from Equation (91) with
the parameters that best fit the data in (b). The stars stand for M-SWNTs, the open bullets stand for type I
S-SWNTs and the filled bullets stand for type II S-SWNTs. Reprinted figure with permission from P.T. Araujo
and A. Jorio, Physical Status Solidi B, 2008, 245, pp. 2201–2204 [335]. Copyright © Wiley–VCH Verlag
GmbH & Co. KGaA.
general dielectric constant κ on Eii . By “general” we mean that κ comprises the screening from
both the tube core electrons and from the tube environment.24 A dt -dependent effective κ value
for the exciton calculation is needed to reproduce the experimental Eii values consistently. This
dependence is important for the physics of quasi -1D and truly 1D materials generally and can be
used in interpreting optical experiments and environment effects for such materials. Environmental
effects are therefore considered further in Sections 5.2.1 to 5.2.3.
to sample for a particular type of SWNT sample, since the electronic structure of a given (n, m)
tube should be the same, these results indicate that the “alcohol-assisted” SWNTs are surrounded
by an environment with a larger κenv value than the “super growth” sample, thus increasing the
effective κ and decreasing Eii [229], which is consistent with Figure 61, discussed below.
The parameters (a, b, c) thus determined are common for all different samples (a, b, c) = (0.8 ±
0.1, 1.6 ± 0.1, 0.4 ± 0.05) so as to both optimize the correlation between κ and (p, dt , lk ), and
to minimize differences between theory and experiment. Here, it should be mentioned that the
variable lk is involved in the κ function because of the screening by the different environments
exp
Figure 60. Black dots show Eii vs. ωRBM results obtained from resonance Raman spectra taken from a
super-growth SWNT sample. The black open circles (semiconducting; S-SWNTs) and the dark-gray stars
(metallic; M-SWNTs) give Eiical calculated for the bright exciton with a dielectric constant κ = 1 [148]. Along
the x axis, the Eiical values are calculated using the relation ωRBM = 227/dt . Due to computer time availability,
only Eii for tubes with dt < 2.5 nm (i.e., ωRBM > 91 cm−1 ) have been calculated. Transition energies EiiS
(i = 1 to 5) denote semiconducting SWNTs and EiiM (i = 1, 2) denote metallic SWNTs. Reprinted figure
with permission from P.T. Araujo et al., Physical Review Letters 103, p. 146802, 2009 [229]. Copyright ©
(2009) by the American Physical Society.
516 R. Saito et al.
Figure 61. The κ function for: (a) SG, (b) AA, and (c) HiPCO samples. (d) Data for the three different types of
samples are plotted on the same figure with a fitted slope Cκ for each sample. (e)All the κ functions collapse on
to a single line after dividing each function by the corresponding C̃κ . The following symbols are used: E11 (◦),
E22 (×), E33 () and E44 (). Black, red, and blue colors, respectively, denote metallic (mod(2n + m, 3) = 0),
semiconductor type I (mod(2n + m, 3) = 1), and type II (mod(2n + m, 3) = 2) SWNTs. Reprinted figure
with permission from A.R.T. Nugraha et al., Applied Physical Letters 97, p. 91905, 2010 [232]. Copyright
© (2010) American Institute of Physics.
which modify the exciton size. However, we will show below that the selection of lk as a variable
for κ is essential for explaining the difference between metallic and semiconducting SWNTs. In
fact, Equation (92) indicates another scaling relation for excitons, similar to the previously reported
scaling law which relates Ebd with dt , κ, and the “effective mass” μ [275]. However, it is found
that the scaling relation involving μ works well only for S-SWNTs and another scaling function
is needed for M-SWNTs. This is because Ebd for an M-SWNT is screened by free electrons even
for a similar effective mass as that for the photo-excited carriers.
In Figure 61, we show a series of results for the κ function [232] for different samples grown
by different methods: (a) super-growth [316], (b) alcohol-assisted CVD [347], and (c) the high
pressure gas-phase decomposition of CO (HiPCO) [348]. For each sample, the κ function is
S S M
successfully unified for lower energy transitions (E11 , E22 , E11 ) and for the higher energy transitions
S S
(E33 , E44 ). Considering lk explicitly is important for describing the environmental effect for both
metallic and semiconducting SWNTs simultaneously, since lk is very different between metallic
and semiconducting SWNTs, even for similar effective mass values because of the screening of
the π electrons. When we consider a SWNT, the κ values for the higher Eii transitions which have
larger lk−1 values are smaller than those for lower Eii (smaller lk−1 ). Thus, lk−1 (the exciton size
in real space) is also smaller for the higher Eii because only a small amount of the electric field
created by an e–h pair will influence the surrounding materials.25
Qualitatively, the origin of the diameter dependence of κ consists of: (1) the diameter-dependent
exciton size and (2) the amount of electric field which goes into the surrounding material. These
two factors are connected to one another and Ando gave an analytic form for an expression
connecting these two factors [349]. The development of an electromagnetic model is needed to
S
fully rationalize Equation (92). Interestingly, the similarity between the κ values found for E22 and
Advances in Physics 517
Figure 62. δEiienv versus dt , scaled by C̃κ . Circles and triangles, respectively, denote AA and HiPCO samples
(see text). Many square symbols on the zero line denote the SG sample which is taken as the standard. The
inset shows differences between experimental (exp) and calculated (cal) Eii values for all samples, showing
good agreement between experiment and the model calculations. Reprinted figure with permission from
A.R.T. Nugraha et al., Applied Physical Letters 97, p. 91905, 2010 [232]. Copyright © (2010) American
Institute of Physics.
M
E11 shows that the difference between metallic and semiconducting tubes is satisfactorily taken
into account by lk using the random phase approximation (RPA) in calculating ε(q) [148,342].
In Figure 61(d), three sets of data contained in Figure 61(a)–(c) are merged, from which we
know that the three plots when taken together depend only on the difference of the slopes, that is Cκ
of Equation (92). Values of Cκ for the SG, AA and HiPCO samples are 0.84, 1.19 and 1.28, respec-
tively. We expect that such differences in the values for Cκ arise from the environmental effects
on the exciton energies. Therefore, we assume that each Cκ value characterizes the environmental
dielectric constant κenv for that particular sample. The SG sample has the largest Eii and hence the
smallest dielectric constant relative to any of the other samples found in the literature [230], and
so we normalize Cκ of the SG sample to be C̃κ (SG) = 1.00 for simplicity. The values of C̃κ for
the other samples can then be determined by taking the ratio of their Cκ values to that for the SG
sample. Thus, C̃κ for the SG, AA, and HiPCO samples becomes 1.00, 1.42, and 1.52, respectively.
When we use the normalized C̃κ , all points collapse on to a single line, as shown in Figure 61(e),
hence giving justification for the use of this assumption. It is interesting to see that all the lines
shown in Figure 61 cross the horizontal axis at κ = 1 at the same point. This point corresponds to
the large diameter limit beyond which the 1D exciton does not exist. The corresponding diameter
is scaled by p and lk .
In Figure 62, we plot the calculated energy shift δEiienv relative to the SG results as a function
of p/dt . The data on the horizontal axis are the data for the SG sample which should be zero from
the definition of δEiienv . The δEiienv for the AA and HiPCO sample are fitted to a function:
' (
2
p p
δEiienv = EiiSG − Eiienv ≡ C̃κ A+B +C , (93)
dt dt
where A, B and C are parameters common to all types of environments and Eiienv is calculated using
the κ function obtained previously. The best fit is found for A = −42.8 meV, B = 46.34 meV · nm
and C = −7.47 meV · nm2 . In the inset to Figure 62, we show the energy difference between
experiment and theory that is obtained by using the κ function (Equation (92)) and we see all the
data points are within 50 meV for a large range of SWNT diameters and energies.
518 R. Saito et al.
exp
Figure 63. Eii vs. ωRBM results obtained for the “super-growth” (bullets) and the “alco-
hol-assisted” (open circles) SWNT samples. Reprinted figure with permission from P.T. Araujo et al.,
Physical Review Letters 103, p. 146802, 2009 [229]. Copyright © (2009) by the American Physical Society.
Figure 64. (a) The totally symmetric G-band eigenvectors for the (8,4) semiconducting SWNT. The atomic
displacements are almost parallel to the circumference. (b) The totally symmetric G-band eigenvector for the
(9,3) metallic SWNT. The atomic displacements are almost parallel to the carbon–carbon bonds. Reprinted
figure with permission from S. Reich et al., Physical Review B 64, p. 195416, 2001 [280]. Copyright ©
(2001) by the American Physical Society.
First-principles calculations have been performed for different SWNTs by Reich et al. [280].
They found that, while for armchair and zigzag SWNTs, which have higher symmetry than chiral
tube structures, the LO and iTO vibrations of the G-band exist [280], such a definition is not
strictly valid for chiral nanotubes, where the atomic vibrations actually depend on the chiral angle.
In Figure 64(a and b) Reich et al. show the mode displacements for one totally symmetric (A1 )
G-band mode in two different chiral SWNTs. For each case, another G-band mode is expected
with the atomic vibrations perpendicular to those indicated in the figure. Note that the atomic
displacements in Figure 64 are almost perfectly aligned along the circumference in the (8,4)
S-SWNT (a), but mostly parallel to the C–C bonds in the (9,3) M-SWNT (b), and in neither
case is the strict iTO definition applicable. These findings by Reich et al. [280] are qualitatively
consistent with other calculations [86], but the quantitative definitions are model-dependent. There
is still controversy about whether the many peaks within this G-band can be assigned to LO and
iTO type mode behavior, as discussed by Piscanec et al. [193].
The Z- and Y - axes are the SWNT axis direction and the photon propagation direction, respec-
tively. The polarization of the incident and scattered light is given as well as the resonance
condition. EG is the G-band phonon energy [1].
photons, respectively, while the incident polarization i and scattered polarization s appear inside
the parenthesis (is). Since the SWNT experiments are usually made by using microscopes in the
back-scattering configuration, pi and ps are usually Y and −Y , so that the simplified notation (is)
can be applied. In this case, four different scattering geometries are possible, and these are labeled
XX, XZ, ZX and ZZ.
The first-order Raman signal from isolated SWNTs can only be seen when the excitation laser
energy is in resonance with a van Hove singularity (VHS) in the JDOS. Chiral SWNTs exhibit CN
symmetry [351,352] and, following the (is) notation introduced above, selection rules imply that:
(1) totally symmetric A phonon modes are observed for the (ZZ) scattering geometry when either
the incident or the scattered photon is in resonance with Eii , and for the (XX) scattering geometry
when either the incident or the scattered photon is in resonance with Ei,i±1 . (2) E1 symmetry
modes can be observed for the (ZX) scattering geometry for resonance of the incident photon with
the Eii VHSs, or for resonance of the scattered photon with the Ei,i±1 VHSs, while for the (XZ)
scattering geometry, for resonance of the incident photon with the Ei,i±1 VHSs, or for resonance of
the scattered photon with the Eii VHSs; (3) E2 symmetry phonon modes can only be observed for
the (XX) scattering geometry for resonance with Ei,i±1 VHSs. Therefore, it is possible to observe
two, four or six G-band peaks, depending on the resonance conditions and on the polarization
scattering geometry, as summarized in Table 4.
Figure 65. (a) Schematic picture of the G-band atomic vibrations along the nanotube circumference and
along the nanotube axis of a zigzag nanotube. (b) The Raman-active modes with A, E1 and E2 symmetries
and the corresponding cutting lines μ = 0, μ = ±1 and μ = ±2 in the unfolded 2D BZ. The points of
the cutting lines are shown by solid dots. Reprinted from Raman Spectroscopy of Carbon Nanotubes, M.S.
Dresselhaus et al. [167]. Copyright © (2005) with permission from Elsevier.
Advances in Physics 521
Figure 66. (a) G-band polarization dependence from one isolated semiconducting SWNT sitting on a Si/SiO2
substrate [357]. Both incident and scattered light are polarized parallel to each other and vary from ZZ (bottom)
to XX (middle) and back to ZZ (top). (b,c) G-band polarization scattering geometry dependence from two
isolated SWNTs. The Lorentzian peak frequencies are given in cm−1 . The incident angles θS and θS between
the light polarization and the SWNT axis directions are not known a priori, but have been assigned as
θS ∼ 0◦ and θS ∼ 90◦ based on the relative intensities of the polarization behavior of the G-band modes and
the expected selection rules and frequencies (see Section 5.3.3) for the A1 , E1 and E2 modes [352]. Reprinted
figure with permission from A. Jorio et al., Physical Review B 65, p. R121402, 2002 [357]. Copyright ©
(2009) by the American Physical Society.
Before showing the experimental results related to the selection rules discussed above, we
introduce a general polarization behavior that is not accounted for in these selection rules and
is responsible for the totally symmetric A1 modes in the G-band spectra which dominate these
spectra. Carbon nanotubes are nano-antennas, and both the absorption and emission of light are
suppressed when the light is polarized perpendicular to the nanotube axis. This phenomenon is
called as the depolarization effect [353,354], where photo-excited carriers screen the electric field
of the cross-polarized light inside the carbon nanotube [234,353,354]. Considering these effects, it
is clear that the Raman intensity is generally largest for the (ZZ) scattering polarization geometry,
while the signal will be suppressed for the (XX) scattering, as shown in Figure 66(a) [355–357].
In addition, remember that the resonance energies are different for light along Z (Eii ) and along
X (Eii±1 ), as described in Table 4. This means that one rarely observes the Raman signals from
parallel (ZZ) and perpendicular ((ZX), (XZ) or (XX)) polarization directions simultaneously.
Now we can discuss experimental results related to the symmetry selection rules for the dif-
ferent scattering geometries [351,357,358]. Figure 66(b) shows the G-band Raman spectra from a
semiconducting SWNT with three different directions for the incident light polarization. Consid-
ering θS the initial angle between light polarization and the nanotube axis, unknown a priori, the
three spectra in Figure 66(b) were acquired with θS , θS + 40◦ and θS + 80◦ . Six well-defined G-
band peaks are observed with different relative intensities for the different polarization geometries
and, based on the selection rules and on their frequencies (see Section 5.3.3), they are assigned
as follows: 1565 and 1591 cm−1 are A1 symmetry modes; 1572 and 1593 cm−1 are E1 symmetry
modes; and 1554 and 1601 cm−1 are E2 symmetry modes. Figure 66(c) shows the G-band Raman
spectra obtained from another semiconducting SWNT, the two spectra with θS and θS + 90◦ ,
522 R. Saito et al.
Figure 67. ωG (open symbols) vs. ωRBM (bottom axis) and vs. 1/dt (top axis) for S-SWNTs.
Explicit experimental G-band data obtained with Elaser = 1.58, 2.41 and 2.54 eV are presented. Solid
symbols connected by solid lines give results obtained from ab initio calculations [249], down-
shifted by about 1% to fit the experimental data [352]. Reprinted figure with permission from
O. Dubay and G. Kresse, Physical Review B 67, p. 35401, 2003 [249]. Copyright © (2003) by the American
Physical Society.
where θS is also unknown a priori. The spectra requires four sharp Lorentzians for a good line-
shape fitting, plus a broad feature at about 1563 cm−1 . This broad feature (FWHM ∼ 50 cm−1 )
is sometimes observed in weakly resonant G-band spectra from SWNTs, and is generated by
defect-induced DR processes. For the sharp peaks, based on the polarization Raman studies and
on their relative frequencies [351], the 1554 and 1600 cm−1 peaks are E2 symmetry modes, while
the 1571 and 1591 cm−1 peaks are unresolved (A1 + E1 ) symmetry modes, with their relative
intensities depending on the incident light polarization direction [351].
Generally speaking, the intensity ratio between the two scattering geometries ZZ : XX can
assume values either larger or smaller than 1, depending on the resonance condition. Observation
of the spectra with (XX) polarization in Figure 66(c) indicates resonance with a Ei,i±1 optical
transition. In samples with a large diameter distribution, e.g. dt from 1.3 nm up to 2.5 nm in
Ref. [351], both the Eii and Ei,i±1 transitions can occur within the resonance window of the same
laser, and an average value of ZZ : XX = 1.00 : 0.25 was then observed. However, for most
isolated SWNTs, the Raman intensities exhibit an intensity ratio IZZ : IXX ∼ 1 : 0. This “antenna
effect” is observed for samples in resonance with only Eii electronic transitions, as shown in
[355–357].
Finally, the above discussion is related to phonon confinement within the first-order single
resonance process [333,350,351]. There has been interesting discussion on whether the features
in the multi-peak G-band could all be assigned to A1 symmetry modes [355,359] originating from
a defect-induced DR Raman scattering process [145]. It seems that both cases, i.e. a multi-peak
feature from multi-symmetry scattering and from an A1 symmetry feature induced by disorder
are possible in very disordered samples [360].
Figure 68. Experimental intensity plot of the G-band spectrum of a metallic SWNT as a function of elec-
trochemical gate potential. For this nanotube the charge neutrality point, corresponding to the Dirac point,
is 1.2 V. Adapted figure with permission from M. Farhat et al., Physical Review Letters 99, p. 145506, 2009
[201]. Copyright © (2009) by the American Physical Society.
in Figure 67. The relation between ωRBM (bottom axis – for experimental results) and inverse
nanotube diameter (top axis – for theoretical results) were made in this figure considering 1/dt =
ωRBM /248. This ωRBM = 248/dt relation [111] was broadly used in the early years of single
nanotube spectroscopy (2001–2005). The present understanding uses Equation (86) with Ce =
0.065, which is applicable to a larger variety of samples.
The spectra in this work were fit using 2, 4 or 6 peaks with FWHM γG ∼ 5 cm−1 , which is the
natural linewidth for the G-band modes [340]. The diameter dependent downshift in frequency
comes from strain and from curvature-induced mixing of low frequency out-of-plane components.
In the time-independent perturbation picture, the ωGLO mode frequency is expected to be indepen-
dent of diameter, since the atomic vibrations are along the tube axis. In contrast, the ωGTO mode
has atomic vibrations along the tube circumference, and increasing the curvature increases the
out-of-plane component, thus decreasing the spring constant with a 1/dt2 dependence. This picture
holds for semiconducting SWNTs, where G+ now stands for the LO mode, and G− stands for
the iTO mode [171]. However, for metallic SWNTs the picture is different: G+ now stands for
the iTO mode, and G− stands for the LO mode [193,249]. The G-band profile in this case is very
different and depends strongly on doping, as shown in Figure 68, and this behavior can only be
understood within a time-dependent perturbation picture. In this section, we focus on the spectra
from semiconducting SWNTs. Metallic SWNTs are discussed in Section 5.4.
The dt dependence of the frequencies for each of the three higher frequency G+ band modes
(A1 , E1 and E2 ) observed experimentally are in very good agreement with theory [249], showing
basically no diameter dependence. For the lower frequency G− band modes, both ab initio calcu-
lations and the experimental results show a considerable dt dependence of the mode frequency.
The experimental G-band frequencies from semiconducting SWNTs can be fit with [352]
C
ωG = 1592 − β
, (94)
dt
524 R. Saito et al.
where the values for the various parameters are: β = 1.4, CA1 = 41.4 cm−1 nm, CE1 =
32.6 cm−1 nm and CE2 = 64.6 cm−1 nm.
A simpler formula for the G-band intensities has also been proposed in Ref. [171], where
the G-band fitting is not performed with six Lorentzians, but rather by considering just the peak
value for the two most intense features. In this case, the higher frequency G+ peak is diameter
independent and the lower frequency G− feature decreases in frequency with a 1/dt2 dependence.
In practice, the G band can be used for a diameter determination for both semiconducting and
metallic SWNTs by using the formula from [171]. For more detailed studies, the more complete
analysis discussed above and including dynamic effects [193] and environmental effects [232]
should be used.
5.4. Kohn Anomaly effect on the G-band and the RBM mode
In this section, the first-order spectra for metallic SWNTs are discussed giving special attention
to the dependence of these spectral features on the Fermi level position and the KA phenomena
that account for the special properties of these spectral features for metallic SWNTs [193]. First,
we discuss the characteristics of the KA for the G-band in Section 5.4.1 through Section 5.4.4,
followed by an elaboration of the KA for the RBM feature in Section 5.4.5.
for semiconducting SWNTs, which means that the G-band phonons are unable to create real e–
h excitations across a typical non-zero bandgap. Therefore, lifetime broadening of the phonon
is not expected, and the G-band peak FWHM should not change significantly. Still, virtual e–h
excitations contribute to the softening of the phonon energies [366]. Both the iTO and LO modes
couple to intermediate e–h pairs, and the iTO mode is expected to show a greater EF -dependent
frequency shift, most significantly in larger diameter nanotubes, as the band gap energy approaches
the phonon energy [366]. This behavior in semiconducting SWNTs is opposite to metallic SWNTs.
Figure 69. (a) Confocal image of the G-band integrated intensity using a laser wavelength λlaser = 632.8 nm
(Elaser = 1.96 eV). The inset shows a schematic view of the SWNT (yellow) lying on top of the miscut quartz.
(b) The G-band Raman spectra obtained at the 41 points indicated and numbered in (a). There is a modulation
of the higher frequency G+ feature (∼1590-1605 cm−1 ) along with the appearance and disappearance of the
lower frequency G− feature (∼1540 cm−1 ). This appearance and disappearance of the G+ and G− features
are related to the tube-substrate morphology and interaction. Adapted with permission from J.S. Soares et al.,
Nono Letters 10, pp. 5043–5048, 2010 [324]. Copyright © (2010) American Chemical Society.
Figure 70. (a) An intermediate e–h pair state that contributes to the energy shift of the optical phonon
modes is depicted. A phonon mode is denoted by a zigzag line and an e–h pair is represented by a loop.
The low energy e–h pair satisfying 0 ≤ E ≤ 2|EF | is forbidden at zero temperature by the Pauli principle.
(b) The energy correction to the phonon energy by an intermediate e–h pair state, especially the sign of
the correction, depends on the energy of the intermediate state as h(E). Adapted figure with permission
from K. Sasaki et al., Physical Review B 78, pp. 235405–235411, 2008 [365]. Copyright © (2008) by the
American Physical Society.
(softening). Thus, when the phonon hardening processes are suppressed by increasing the Fermi
energy, we get a large phonon softening and broadening at 2EF = ±0.2 eV. This is the origin of
the observed W lineshape [196]. Recent measurements of the G-band spectra at T = 12 K show
phonon anomalies at EF = ±ωG /2 that could be clearly distinguished experimentally [200].
Furthermore, the el–ph matrix element for this virtual excitation has a chiral angle dependence
that acts differently for iTO and LO phonons [196]. For example, the el–ph interaction for the
iTO phonon is absent (no KA effect) for armchair nanotube, while phonon hardening of the iTO
mode occurs for zigzag nanotubes. Thus, a detailed analysis of the LO and iTO phonon modes as a
function of the Fermi energy gives important information about the (n, m) assignment for metallic
carbon nanotubes. For a detailed description of these phenomena, see [193,196,205,246].
Advances in Physics 527
Figure 71. Deviation of the experimentally observed RBM frequency (ωRBM ) from the linear depen-
dence given by (218.3/dt + 15.9), as a function of θ for a particular HiPCO nanotube sample.
Filled, open and crossed circles denote M-SWNTs, type I and type II S-SWNTs, respectively. The
dotted lines show an experimental accuracy of ±1 cm−1 . Reprinted figure with permission from
A. Jorio et al., Physical Review B 71, p. 75401, 2005 [369]. Copyright © (2005) by the American Physical
Society.
Figure 72. (a) 2D plot of the Elaser dependence for the Raman spectra of SWNT bundles in the intermedi-
ate frequency mode (IFM) range. The bright, light areas indicate high Raman scattering intensity. Arrows
point to five well-defined ωIFM features. (b) Raman spectra in the corresponding IFM range are taken
at discrete laser excitation energies Elaser = 2.05, 2.20, 2.34, and 2.54 eV. Reprinted figure with permission
from C. Fatini et al., Physical Review Letters 93, p. 87401, 2005 [371]. Copyright © (2003) by the American
Physical Society.
Figure 73. (a) G -band data for ωG for a SWNT bundle sample taken from [222] after subtracting the
linear dispersion 2420 + 1106Elaser from the ωG vs. Elaser data shown in the inset. (b) Optical transition
energies Eii as a function of diameter for SWNTs. The vertical lines denote the diameter range of the SWNT
bundle used in the G -band dispersion experiment shown in (a). Reprinted figure with permission from
A.G. Souza Filho et al., Physical Review B 65, p. 35404, 2001 [376]. Copyright © (2001) by the American
Physical Society.
which is consistent with observations in graphene and graphite. However, different from graphene
and graphite, the G -band dispersion in SWNTs exhibits a superimposed oscillatory behavior as a
function of Elaser , as shown in Figure 73(a), where the linear dispersion effect was subtracted from
the experimentally observed frequencies. The oscillatory behavior seen in Figure 73(a) is due to
the ωG dependence on tube diameter, as discussed below.
The G -band frequency (ωG ) depends on tube diameter (dt ) because of a force constant soft-
ening, which is associated with the curvature of the nanotube wall. Experiments on isolated tubes
530 R. Saito et al.
Figure 74. Cutting lines for two metallic SWNTs, one zigzag and one armchair, in the unfolded 2D BZ of
graphene. The wavevectors ki point with arrows to the locations where the van Hove singularities occur.
Reprinted figure with permission from G.G. Samsonidze et al., Physical Review Letters 90, p. 27403, 2003
[378]. Copyright © (2003) by the American Physical Society.
Figure 75. The G -band Raman spectra for (a) a semiconducting (15, 7) and (b) a metallic (27, 3) SWNT,
showing two-peak structures [375,378,379], respectively. The vicinity of the K point in the unfolded BZ is
shown in the lower part of the figure, where the equi-energy contours for the incident Elaser = 2.41 eV and the
scattered Elaser − EG = 2.08 eV photons, together with the cutting lines and wave vectors for the resonant van
S = 2.19 eV, E S = 2.51 eV, E M(L) = 2.04 eV, E M(U) = 2.31 eV) are shown. Reprinted
Hove singularities (E33 44 22 22
figures from A.G. Souza Filho et al., Physical Review B 65, p. 85417, 2002 [375]. Copyright © (2002) by
the American Physical Society and A.G. Souza Filho et al., Chemical Physics Letters 354, pp. 62–68,
Copyright © (2002) with permission from Elsevier.
an equi-energy contour, thus causing a chiral angle dependence on the ki value where a particular
excited state ki occurs. The states at ki are those responsible for the dominant optical spectra
observed in SWNTs, including the DR features. The presence of cutting lines in carbon nanotubes
is expected to affect all the dispersive Raman features [376], but here we focus on the G -band,
because the G -band dispersion is very large and is an interesting effect.
The two-peak G -band Raman features in the Raman spectra observed from semiconducting
and metallic isolated nanotubes are shown in Figure 75(a) and (b), respectively. The presence of
two peaks in the G -band Raman feature indicates the resonance with both the incident Elaser and
scattered Elaser –EG photons, respectively, with two different van Hove singularities (VHSs) for
the same nanotube. Elaser and Elaser –EG are defined in Figure 75(a) and (b) below the G -band
spectra, by the outer and inner equi-energy contours near the 2D BZ boundary, in which the cutting
lines are shown and the trigonal warping of these constant energy contours can be seen [380]. The
two peaks in Figure 75(a) and (b) can be associated with the phonon modes of the wave vectors
qi = −2ki , where i = 3, 4, 2L, 2U relate to E33
S S
, E44 ML
, E22 MU
and E22 , respectively, and the electronic
wave vectors ki are shown in the lower part of Figure 75. For the semiconducting SWNT shown
in Figure 75(a), the resonant wave vectors k3 and k4 have different magnitudes, k4 − k3 K1 /3,
resulting in twice the difference for the phonon wave vectors, q4 − q3 2K1 /3 = 4dt /3, so that
the splitting of the G -band Raman feature arises from the phonon dispersion ωph (q) around the
K point. In contrast, for the metallic nanotube (M-SWNT) shown in Figure 75(b), the resonant
wave vectors k2L and k2U have roughly equal magnitudes and opposite directions away from the K
532 R. Saito et al.
Figure 76. Localized excitonic emission in a semiconducting SWNT. (a) Photoluminescence emission at
λem = 900 nm from a single SWNT. (b) Raman spectrum recorded from the same SWNT. The spectral
position of the RBM, ωRBM = 302 cm−1 , together with the λem = 900 nm information, leads to the (9,1)
assignment for this tube. (c) Near-field photoluminescence image of the SWNT revealing localized excitonic
emission. (d)–(e) Near-field Raman imaging of the same SWNT, where the image contrast is provided by
spectral integration over the G and D bands, respectively. (f) Corresponding topography image. The circles
indicate localized photoluminescence (c) and defect-induced (D band) Raman scattering (e). The scale bar
in (c) denotes 250 nm. (g) Evolution of the G -band spectra near the defective segment of the (9,1) SWNT.
The spectra were taken in steps of 25 nm along the nanotube, showing the defect-induced G peak (dotted
Lorentzian). The asterisks denote the spatial locations where localized photoluminescence and defect-induced
D-band Raman scattering were measured (see circles in (c) and (e), respectively). Reprinted with permission
from I.O. Maciel et al., Nature Materials 7, pp. 878–883, 2008 [293]. Copyright © (2008) American Institute
of Physics.
point, so that the splitting of the G -band Raman feature for metallic nanotubes arises from the
anisotropy of the phonon dispersion ωph (q) around the K point [378], which we identify with the
phonon trigonal warping effect. Overall, the presence of two peaks in the DR Raman features of
isolated carbon nanotubes is associated with quantum confinement effects expressed in terms of
cutting lines.
Of course the G -band is not the only feature to exhibit an (n, m) dependence for SWNT
systems. Actually, all the DR features are expected to exhibit such a chirality dependence. The
stronger the dispersive behavior, the larger is the (n, m) chirality dependence. For this reason,
SWNTs with smaller dt show larger frequency splittings and larger G -band shift effects. The
(n, m) dependence of other combination modes, such as the iTO+LA combination mode near the
point, have also been studied in some detail (see e.g. Ref. [227]).
Advances in Physics 533
of our basic knowledge. For example, our understanding related to the intensity of the DR features
is marginal. Also, there is still controversy about the assignment of the 2450 cm−1 peak, basically
because different DR features fall at this frequency, and it is not clear which of them is responsible
for the peak. Of course, the relative intensity for the features related to defects depend on the
number of defects. However, the relative intensity of the DR features (even those unrelated to
defects) change from sample to sample. Graphene, as the prototype material, may shed light into
this and other basic issues related to inelastic light scattering.
and (3) the observation of previously elusive transitions from metallic nanotubes [230]. However,
it is not yet clear why this sample is special. Clarifying such a result could help in developing an
effective standard reference material in the graphene area.
the development of new theories to explain the newly found effects. Raman spectroscopy thus
helps the development of nanotechnology in its most basic sense, and allows study of physical
processes that are expected to occur at the nanometer level. In following the words of Professor
Richard Feynman in his lecture on 29 December 1959, that there is plenty of room for discovery
at the bottom, below 10 nm spatial resolution, below 1 meV energy resolution and below 10 fs
time resolution, and for the operation of materials at more than 1 THz frequency, 1000 T magnetic
field, and 1 TPa pressure.
Acknowledgements
The MIT authors acknowledge the support under NSF Grant DMR 10-04147. A.J. acknowledges
the financial support from the Brazilian agencies CNPq, CAPES and FAPEMIG. R.S. acknowl-
edges a Grant-in-Aid (No. 20241023) from the MEXT, Japan. We would like to thank the referee
for his/her suggestions which improved the quality of this publication.
For all figures reprinted from American Physical Society material readers may view, browse,
and/or download material for temporary copying purposes only, provided these uses are for
non-commercial personal purposes. Except as provided by law, this material may not be fur-
ther reproduced, distributed, transmitted, modified, adapted, performed, displayed, published, or
sold in whole or part, without prior written permission from the American Physical Society.
Notes
1. The B atom in Figure 1 gives a brighter STM images than the A atom, since there are electronic energy
bands for the B atom near the Fermi energy.
2. The G peak in the Raman spectra of sp2 carbons is often called the 2D peak. It should be noted that
the mechanisms involved in the 2D double resonance (DR) processes are different from those for the G
peak which involves only two phonons. The G peak involves only two K point phonons, whereas the 2D
feature arises from the second-order D that includes a DR D-band process involving a K point phonon
and an elastic scattering process. In this review, we distinguish between the G and the 2D scattering
process and the actual values of their frequencies.
3. A Stokes process is a terminology used to denote the loss of photon energy in a scattering process. Here
the Stokes photoluminescence process is independent of the Stokes Raman process.
4. Time-dependent perturbation theory tells us that the amplitude of the wavefunctions for excited states
oscillates as a function of time if the photon energy does not match the excited-state energy, which is
the physical meaning of a virtual transition.
5. A notch filter is an optical filter that suppresses a specified range of energies of the incident light.
6. By connecting monochromators in a serial way, the resolution of a monochromator is significantly
improved although the actual signal becomes increasingly weak. Furthermore, extra monochromators
can be used as an energy-tunable filter to reject the elastically scattered light in contrast to the notch
filters, which are wavelength-specific.
7. The solution of a forced damped harmonic oscillator is not generally a Lorentzian lineshape but
approaches the Lorentzian function for ωq q . However, if ωq approaches q , the lineshape departs
from a Lorentzian function.
8. The designation π of the π band comes from its value of angular momentum which is 1.
9. Here mod denotes an integer function for evaluating the modulus for an (n, m) SWNT where we use the
notation mod (6, 3) = 0, mod (7, 3) = 1 and mod (8, 3) = 2 as an illustrative example. Some authors
use Mod 1 and Mod 2 to denote a semiconducting nanotube depending on mod(n − m, 3) = 1 or 2.
There is thus a one-to-one correspondence between type I (II) and Mod 2 (1) semiconducting nanotubes
appearing in the literature, and this is clarified in Figure 11.
10. It should be noted that there is a logarithmic 2D van Hove singularity in the density of states of graphene
at the saddle point of the energy band near the M points (center of the hexagonal edges) of the BZ.
538 R. Saito et al.
11. In time-dependent perturbation theory, the mixing (or transition) of the excited states occurs as a function
of time with some finite and often measurable probability. The virtual states are defined by such a linear
combination of excited states with some probability. The probability for the occupation of an excited
state can be large when the energy difference between the excited states and the energy of the external
field is relatively small. In such a case, we can say that the transition is resonant with the excited state.
12. Here “real absorption” means that the photo-excited electron can be in the excited states for a sufficient
time, for example, 1 ns, so that the electron can be probed in the excited state. A material can scatter
photons in a virtual process.
13. It is noted that not all even (odd) vibration modes under inversion symmetry are Raman (IR)-active
modes.
14. The two-phonon process involving one-phonon emission and one-phonon absorption does not contribute
to the Raman spectra but rather gives a correction to the effective Rayleigh spectral process.
15. Here q is the real phonon wavevector, measured from the point, while in defining qDR , the k and k
vectors are measured from the K point or alternatively, with respect to the K point.
16. It is only when crystalline disorder is present that the first-order q = 0 phonons can be observed, as
discussed in Section 4.3.
17. From the matrix element α p , we can deduce another matrix element,
βp (τ ) = φμ (r)∇v(r − τ)φν (r − τ) dr
(45a)
= φν (r)∇v(r)φμ (r + τ) dr = βp (τ )Î(βp ).
However, the integral in Equation (45a) can also √ be expressed by α terms [268].
18. The phonon amplitude is proportional to Aν (q) n̄ν (q) in which the temperature dependence of the
amplitude is expressed by n̄ν (q) given by Equation (47).
19. It is noted that the minus sign corresponds to a symmetric wavefunction and that the plus sign corresponds
to an anti-symmetric wavefunction.
20. A virtual state is a linear combination of real states. When a virtual state is close to an exciton state, the
virtual state contains a large component of the exciton states. This is the reason for the approximation
used in obtaining a representation for the virtual state.
21. A cutting line is defined by the 1D BZ of an SWNT in the 2D BZ of graphene [32,135,136].
22. Other formula for ωRBM can be used here, too. The difference is within 1–3 cm−1 .
23. The subscript htt in Thtt denotes heat treatment temperature.
24. Here core electrons refer to 1s and σ electrons. The screening by π electrons is independently considered
by the polarization function within the RPA (random phase approximation) [120,148]. See Section 5.2.1
for further details.
25. The Bethe–Salpeter equation is independently solved for each value of κ. When we obtain Eii values as
a function of κ by solving the Bethe–Salpeter equation many times, then Eii with different i values are
adapted from the different κ value. Since the Eii eigenvalues come from different cutting lines, there is
no problem with the orthogonality of the wavefunctions.
References
[1] A. Jorio, R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Raman Spectroscopy in Graphene Related
Systems, Wiley-VCH Verlag GmbH & Co KGaA, Weinheim, Germany, 2010.
[2] B.T. Kelly, Physics of Graphite, Applied Science Publishers, London and New Jersey, 1981, p. 477.
[3] M.M. Lucchese, F. Stavale, E.H. Martins Ferreira, C. Vilani, M.V.O. Moutinho, R.B. Capaz, C.A.
Achete, and A. Jorio, Carbon 48(5) (2010), pp. 1592–1597.
[4] L.M. Malard, M.H.D. Guimarães, D.L. Mafra, M.S.C. Mazzoni, and A. Jorio, Phys. Rev. B 79, (2009)
p. 125426.
[5] L.M. Malard, M.A. Pimenta, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rep. 473 (2009), pp. 51–87.
[6] R.W.G. Wyckoff, Crystal Structures, Wiley, 1963. Vol 6.
Advances in Physics 539
[7] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V. Grigorieva, and
A.A. Firsov, Science 306(5696) (2004), p. 666.
[8] A.K. Geim and A.H. MacDonald, Phys. Today 60(35) (2007), p. 35.
[9] H. Lipson and A.R. Stokes, Nature 149(3777) (1942), p. 328.
[10] A.W. Moore, Highly Oriented Pyrolytic Graphite, Vol. 11. Marcel Dekker, New York, 1973.
[11] J. Heremans, C.H. Olk, G.L. Eesley, J. Steinbeck, and G. Dresselhaus, Phys. Rev. Lett. 60 (1988),
p. 452,
[12] A.W. Moore, Nature 221(5186) (1969), pp. 1133–1134.
[13] A.W. Moore, Chemistry and Physics of Carbon, Vol. 17. Marcel Dekker, New York, 1981.
[14] A.W. Moore, A.R. Ubbelohde, and D.A. Young, Br. J. Appl. Phys. 13 (1962), p. 393.
[15] A.R. Ubbelohde, Carbon 7(5) (1969), pp. 523–530.
[16] I.L. Spain, A.R. Ubbelohde, and D.A. Young, Phil. Trans. R. Soc. London Ser. A, Math. Phys. Sci.
262(1128) (1967), pp. 345–386.
[17] R. Bacon, J. Appl. Phys. 31(2) (1960), p. 283.
[18] M. Endo, T. Koyama, and Y. Hishiyama, Jpn. J. Appl. Phys. 15 (1976), pp. 2073–2076.
[19] M.S. Dresselhaus, G. Dresselhaus, and K. Sugihara, Graphite Fibers and Filaments 5. Springer Series
in Materials Science, Springer, Berlin, 1988.
[20] M. Endo, M.S. Strano, and P.M. Ajayan, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008.
[21] A. Jorio, M.S. Dresselhaus, and G. Dresselhaus, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008.
[22] H.W. Kroto, J.R. Heath, S.C. O’Brien, R.F. Curl, and R.E. Smalley, Nature 318(6042) (1985),
pp. 162–163.
[23] M.S. Dresselhaus, G. Dresselhaus, and P.C. Eklund, Science of Fullerenes and Carbon Nanotubes,
Academic Press, New York, NY/San Diego, CA, 1996.
[24] S. Iijima, Nature 354 (1991), pp. 56–58.
[25] A. Oberlin, M. Endo, and T. Koyama, Carbon 14(2) (1976), pp. 133–135.
[26] A. Oberlin and T. Endo, J. Cryst. Growth 32(3) (1976), pp. 335–349.
[27] A. Oberlin, Carbon 22 (1984), p. 521.
[28] L.V. Radushkevich and V.M. Lukyanovich, Zurn Fisic Chim. 26 (1952), pp. 88–95.
[29] M. Monthioux and V.L. Kuznetsov, Carbon 44 (2006), pp. 1621–1623.
[30] S. Iijima and T. Ichihashi, Nature 363(6430) (1993), pp. 603–605.
[31] D.S. Bethune, C.H. Klang, M.S. De Vries, G. Gorman, R. Savoy, J. Vazquez, and R. Beyers, Nature
363 (1993), p. 605.
[32] R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Physical Properties of Carbon Nanotubes, Imperial
College Press, London, 1998. London.
[33] S. Reich, C. Thomsen, and P. Ordejon, Elastic Properties and Pressure-induced Phase Transitions of
Single-walled Carbon Nanotubes, Vol. 235, Wiley Online Library, 2003.
[34] E. Joselevich, H. Dai, J. Liu, K. Hata, and A.H. Windle, Carbon Nanotubes: Advanced Topics in the
Synthesis, Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 101–164.
[35] M.S. Arnold, A.A. Green, J.F. Hulvat, S.I. Stupp, and M.C. Hersam, Nat. Nanotechnol. 1(1) (2006),
pp. 60–65.
[36] M. Terrones, A.G. Souza Filho, and A.M. Rao, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 531–566.
[37] M. Yudasaka, S. Iijima, and V.H. Crespi, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 605–629.
[38] B.I. Yakobson, Appl. Phys. Lett. 72(8) (1998), p. 918.
[39] C.D. Spataru, S. Ismail-Beigi, R. Capaz, and S.G. Louie, Carbon Nanotubes: Advanced Topics in the
Synthesis, Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 195–228.
[40] T. Ando, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure, Properties and Applications,
Vol. 111, Springer, Berlin, 2008, pp. 229–250.
[41] R. Saito, C. Fantini, and J. Jiang, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure,
Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 251–286.
540 R. Saito et al.
[42] J. Lefebvre, S. Maruyama, and P. Finnie, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 287–319.
[43] Y.-Z. Ma, T. Hertel, Z.V. Vardeny, G.R. Fleming, and L. Valkunas, Carbon Nanotubes: Advanced
Topics in the Synthesis, Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008,
pp. 321–352.
[44] T.F. Heinz, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure, Properties and
Applications, Vol. 111, Springer, Berlin, 2008, pp. 353–369.
[45] A. Hartschuh, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure, Properties and
Applications, Vol. 111, Springer, Berlin, 2008, pp. 371–392.
[46] J. Kono, J.R. Nicholas, and S. Roche, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure,
Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 393–421.
[47] P. Avouris, M. Freitag, and V. Perebeinos, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 423–454.
[48] J. Wu, W. Walukiewicz, W. Shan, E. Bourret-Courchesne, J.W. Ager III, K.M. Yu, E.E. Haller,
K. Kissell, S.M. Bachilo, and R.B. Weisman, Phys. Rev. Lett. 93(1) (2004), p. 17404.
[49] M.S. Biercuk, S. Ilani, and C.M. Marcus, Carbon Nanotubes: Advanced Topics in the Synthesis,
Structure, Properties and Applications, Vol. 111, Springer, Berlin, 2008, pp. 455–492.
[50] L. Kavan and L. Dunsch, Carbon Nanotubes: Advanced Topics in the Synthesis, Structure, Properties
and Applications, Vol. 111, Springer, Berlin, 2008, pp. 567–603.
[51] M. Kalbac, H. Farhat, L. Kavan, J. Kong, and M.S. Dresselhaus, Nano Lett. 8(10) (2008), pp. 3532–
3537.
[52] P.R. Wallace, Phys. Rev. 71(9) (1947), pp. 622–634.
[53] H.P. Boehm, A. Clauss, U. Hofmann, and G.O. Fischer, Z. Naturforschung B 17(3) (1962), p. 150.
[54] A.M. Affoune, B.L.V. Prasad, H. Sato, T. Enoki, Y. Kaburagi, and Y. Hishiyama, Chem. Phys. Lett.
348(1–2) (2001), pp. 17–20.
[55] C. Berger, Z. Song, T. Li, X. Li, A.Y. Ogbazghi, R. Feng, Z. Dai, A.N. Marchenkov, E.H. Conrad, P.N.
First, and W.A. de Heer, J. Phys. Chem. B 108(52) (2004), pp. 19912–19916.
[56] T. Enoki,Y. Kobayashi, C. Katsuyama,V.Y. Osipov, M.V. Baidakova, K. Takai, K.I. Fukui, andA.Y.Vul,
Diam. Relat. Mater. 16(12) (2007), pp. 2029–2034.
[57] Y. Kobayashi, K. Fukui, T. Enoki, K. Kusakabe, andY. Kaburagi, Phys. Rev. B 71(19) (2005), p. 193406.
[58] Y. Kobayashi, K. Fukui, T. Enoki, and K. Kusakabe, Phys. Rev. B 73(12) (2006), p. 125415.
[59] L.G. Cancado, M.A. Pimenta, B.R.A. Neves, G. Medeiros-Ribeiro, T. Enoki, Y. Kobayashi, K. Takai,
K. Fukui, M.S. Dresselhaus, and R. Saito, Phys. Rev. Lett. 93(4) (2004), p. 47403.
[60] A.K. Geim and K.S. Novoselov, Nat. Mater. 6(3) (2007), pp. 183–191.
[61] A. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, and A.K. Geim, Rev. Mod. Phys. 81(1)
(2009), pp. 109–162.
[62] M.S. Dresselhaus and P.T. Araujo, ACS Nano 4(11) (2010), pp. 6297–6302.
[63] C. Lee, X. Wei, J.W. Kysar, and J. Hone, Science 321(5887) (2008), p. 385.
[64] A.A. Balandin and O.L. Lazarenkova, Appl. Phys. Lett. 82 (2003), p. 415.
[65] D. Teweldebrhan and A.A. Balandin, Appl. Phys. Lett. 94 (2009), p. 13101.
[66] K.I. Bolotin, K.J. Sikes, Z. Jiang, M. Klima, G. Fudenberg, J. Hone, P. Kim, and H.L. Stormer, Solid
State Commun. 146(9–10) (2008), pp. 351–355.
[67] S. Morozov, K. Novoselov, M. Katsnelson, F. Schedin, D. Elias, J. Jaszczak, and A. Geim, Phys. Rev.
Lett. 100(1) (2008), p. 16602.
[68] B. Lee, Y. Chen, F. Duerr, D. Mastrogiovanni, E. Garfunkel, E.Y. Andrei, and V. Podzorov, Nano Lett.
10 (2010), pp. 1407–1433.
[69] X. Du, I. Skachko, A. Barker, and E.Y. Andrei, Nat. Nanotechnol. 3(8) (2008), pp. 491–495.
[70] G. Li and E.Y. Andrei, Nat. Phys. 3 (2007), p. 623.
[71] X. Du, I. Skachko, F. Duerr, A. Luican, and E.Y. Andrei, Nature 462 (2009), pp. 192–195.
[72] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, M.I. Katsnelson, I.V. Grigorieva, S.V. Dubonos,
and A.A. Firsov, Nature 438(7065) (2005), pp. 197–200.
[73] K.S. Novoselov, E. McCann, S.V. Morozov, V.I. Fal’ko, M.I. Katsnelson, U. Zeitler, D. Jiang,
F. Schedin, and A.K. Geim, Nat. Phys. 2(3) (2006), pp. 177–180.
Advances in Physics 541
[74] R.R. Nair, P. Blake, A.N. Grigorenko, K.S. Novoselov, T.J. Booth, T. Stauber, N.M.R. Peres, and A.K.
Geim, Science (New York) 320(5881) (2008), p. 1308.
[75] A.F. Young and P. Kim, Nat. Phys. 5(3) (2009), pp. 222–226.
[76] C.W.J. Beenakker, Rev. Mod. Phys. 80(4) (2008), pp. 1337–1354.
[77] M.I. Katsnelson, K.S. Novoselov, and A.K. Geim, Nat. Phys. 2(9) (2006), pp. 620–625.
[78] V.V. Cheianov, V. Fal’ko, and B.L. Altshuler, Science 315(5816) (2007), pp. 1252–1255.
[79] J.M. Pereira Jr., V. Mlinar, F.M. Peeters, and P. Vasilopoulos, Phys. Rev. B 74(4) (2006), p. 45424.
[80] V. Cheianov and V. Fal’ko, Phys. Rev. B 74(4) (2006), p. 041403.
[81] C.W.J. Beenakker, Phys. Rev. Lett. 97(6) (2006), p. 67007.
[82] F. Miao, S. Wijeratne, Y. Zhang, U.C. Coskun, W. Bao, and C.N. Lau, Science 317(5844) (2007),
pp. 1530–1533.
[83] A. Ossipov, M. Titov, and C. Beenakker, Phys. Rev. B 75(24) (2007), p. 241401.
[84] C. Beenakker, A. Akhmerov, P. Recher, and J. Tworzydlo, Phys. Rev. B 77(7) (2008), p. 075409.
[85] T. Pedersen, C. Flindt, J. Pedersen, N. Mortensen, A.-P. Jauho, and K. Pedersen, Phys. Rev. Lett.
100(13) (2008), p. 136804.
[86] J.S. Park, K. Sasaki, R. Saito, W. Izumida, M. Kalbac, H. Farhat, G. Dresselhaus, and M.S. Dresselhaus,
Phys. Rev. B 80(8) (2009), p. 81402.
[87] D.C. Elias, R.R. Nair, T.M.G. Mohiuddin, S.V. Morozov, P. Blake, M.P. Halsall, A.C. Ferrari,
D.W. Boukhvalov, M.I. Katsnelson, and A.K. Geim, Science 323(5914) (2009), p. 610.
[88] A.K. Geim, Science (New York) 324(5934) (2009), pp. 1530–1534.
[89] C. Berger, Z. Song, X. Li, X. Wu, N. Brown, C. Naud, D. Mayou, T. Li, J. Hass, A.N. Marchenkov,
E.H. Conrad, P.N. First, and W.A. de Heer, Electronic confinement and coherence in patterned epitaxial
graphene, Since 312, 1191–1196 (2006).
[90] X. Li, X. Wang, L. Zhang, S. Lee, and H. Dai, Science 319(5867) (2008), p. 1229.
[91] X. Yang, X. Dou, A. Rouhanipour, L. Zhi, H.J. Räder, and K. Müllen, J. Am. Chem. Soc. 130(13)
(2008), pp. 4216–4217.
[92] D.V. Kosynkin, A.L. Higginbotham, A. Sinitskii, J.R. Lomeda, A. Dimiev, B.K. Price, and J.M. Tour,
Nature 458(7240) (2009), pp. 872–876.
[93] L. Jiao, L. Zhang, X. Wang, G. Diankov, and H. Dai, Nature 458(7240) (2009), pp. 877–880.
[94] X. Jia, M. Hofmann, V. Meunier, B.G. Sumpter, J. Campos-Delgado, J.M. Romo-Herrera, H.
Son, Y.-P. Hsieh, A. Reina, J. Kong, M. Terrones, and M.S. Dresselhaus, Science 323 (2009),
pp. 1701–1705.
[95] M. Cardona, in Light Scattering in Solids II, M. Cardona and G. Güntherodt, eds., Vol. 50, Springer,
Berlin, 1982, pp. 19–176.
[96] M.S. Dresselhaus, A. Jorio, M. Hofmann, G. Dresselhaus, and R. Saito, Nano Lett. 10 (2010),
pp. 953–973.
[97] A.C. Ferrari, Raman spectroscopy of graphene and graphite: disorder, electron–phonon coupling,
doping and nonadiabatic effects, July 2007. Solid State Communications, 143 (2007), pp. 47–57.
[98] D. Haberer, D.V. Vyalikh, S. Taioli, B. Dora, M. Farjam, J. Fink, D. Marchenko, T. Pichler, K. Ziegler,
S. Simonucci, M.S. Dresselhaus, M. Knupfer, B. Büchner, and A. Grüneis, Nano Lett. 10 (2010),
pp. 3360–3366.
[99] M.S. Dresselhaus, G. Dresselhaus, and R. Saito, Phys. Rev. B 45(11) (1992), pp. 6234–6242.
[100] R. Saito, T. Takeya, T. Kimura, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 57(7) (1998),
pp. 4145–4153.
[101] E.B. Barros, N. Demir, A.G. Souza Filho, J. Mendes Filho, A. Jorio, G. Dresselhaus, and
M.S. Dresselhaus, Phys. Rev. B 71(16), 2005.
[102] F. Tuinstra and J.L. Koenig, J. Chem. Phys. 53(3) (1970), p. 1126.
[103] A.M. Rao, E. Richter, S. Bandow, B. Chase, P.C. Eklund, K.A. Williams, S. Fang, K.R. Subbaswamy,
M. Menon, A. Thess, R.E. Smalley, G. Dresselhaus, and M.S. Dresselhaus, Science 275(5297) (1997),
pp. 187–191.
[104] C. Thomsen and S. Reich, Phys. Rev. Lett. 85(24) (2000), pp. 5214–5217.
[105] L. Alvarez, A. Righi, T. Guillard, S. Rols, E. Anglaret, D. Laplaze, and J.L. Sauvajol, Chem. Phys.
Lett. 316(3–4) (2000), pp. 186–190.
542 R. Saito et al.
[106] H. Kuzmany, W. Plank, M. Hulman, Ch. Kramberger, A. Grüneis, Th. Pichler, H. Peterlik, H. Kataura,
and Y. Achiba, Eur. Phys. J. B 22(3) (2001), pp. 307–320.
[107] H. Telg, J. Maultzsch, S. Reich, F. Hennrich, and C. Thomsen, Phys. Rev. Lett. 93 (2004), p. 177401.
[108] J. Meyer, M. Paillet, T. Michel, A. Moréac, A. Neumann, G. Duesberg, S. Roth, and J.-L. Sauvajol,
Phys. Rev. Lett. 95(21) (2005), p. 217401.
[109] M. Milnera, J. Kürti, M. Hulman, and H. Kuzmany, Phys. Rev. Lett. 84(6) (2000), pp. 1324–1327.
[110] S. Lefrant, J.P. Buisson, J. Schreiber, O. Chauvet, M. Baibarac, and I. Baltog, Synth. Met. 139(3)
(2003), pp. 783–785.
[111] A. Jorio, R. Saito, J. Hafner, C. Lieber, M. Hunter, T. McClure, G. Dresselhaus, and M.S. Dresselhaus,
Phys. Rev. Lett. 86(6) (2001), pp. 1118–1121.
[112] A.C. Ferrari, J. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D. Jiang,
K. Novoselov, S. Roth, and A. Geim, Phys. Rev. Lett. 97(18) (2006), p. 187401.
[113] D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen, C. Hierold, and L. Wirtz, Nano Lett. 7(2)
(2007), pp. 238–242.
[114] A. Gupta, G. Chen, P. Joshi, S. Tadigadapa, and P.C. Eklund, Nano Lett. 6(12) (2006), pp. 2667–2673.
[115] K. Sasaki, R. Saito, M.S. Dresselhaus, K. Wakabayashi, T. Enoki, New J. Phys. 12 (2010), p. 103015.
[116] C.V. Raman and K.S. Krishnan, Nature 121(3048) (1928), p. 501.
[117] G. Landsberg and L. Mandelstam, Naturwissenschaften 16 (1928), pp. 772–772.
[118] F. Bassani, G. Pastori Parravicini, and R.A. Ballinger, Electronic States and Optical Transitions in
Solids, Pergamon Press, New York, 1975.
[119] N.W. Ashcroft and N.D. Mermin, Solid State Physics, Holt, Rinehart and Winston, New York, 1976.
[120] T. Ando, J. Phys. Soc. Jpn. 66(4) (1997), pp. 1066–1073.
[121] F. Wang, G. Dukovic, L.E. Brus, and T.F. Heinz, Science 308(5723) (2005), p. 838.
[122] J. Maultzsch, R. Pomraenke, S. Reich, E. Chang, D. Prezzi, A. Ruini, E. Molinari, M.S. Strano,
C. Thomsen, and C. Lienau, Phys. Rev. B 72(24) (2005), p. 241402.
[123] S.Y. Zhou, G.-H. Gweon, J. Graf, A.V. Fedorov, C.D. Spataru, R.D. Diehl, Y. Kopelevich, D.-H. Lee,
S.G. Louie, and A. Lanzara, Nat. Phys. 2(9) (2006), pp. 595–599.
[124] A. Bostwick, T. Ohta, T. Seyller, K. Horn, and E. Rotenberg, Nat. Phys. 3(1) (2006), pp. 36–40.
[125] A. Grüneis, C. Attaccalite, T. Pichler, V. Zabolotnyy, H. Shiozawa, S. Molodtsov, D. Inosov,
A. Koitzsch, M. Knupfer, J. Schiessling, R. Follath, R. Weber, P. Rudolf, L. Wirtz, and A. Rubio,
Phys. Rev. Lett. 100(3) (2008), p. 37601.
[126] K. Sugawara, T. Sato, S. Souma, T. Takahashi, and H. Suematsu, Phys. Rev. Lett. 98(3) (2007),
p. 36801.
[127] W. Kohn, Phys. Rev. Lett. 2 (1959), p. 393.
[128] S.D.M. Brown, A. Jorio, P. Corio, M.S. Dresselhaus, G. Dresselhaus, R. Saito, and K. Kneipp, Phys.
Rev. B 63(15) (2001), p. 155414.
[129] U. Fano, Phys. Rev. 124(6) (1961), pp. 1866–1878.
[130] E.H. Martins Ferreira, M.V.O. Moutinho, F. Stavale, M.M. Lucchese, R.B. Capaz, C.A. Achete, and
A. Jorio, Phys. Rev. B 82(12) (2010), p. 125429.
[131] A.C. Ferrari and J. Robertson, Phys. Rev. B 61(20) (2000), pp. 14095–14107.
[132] R.J. Nemanich, S.A. Solin, and R.M. Martin, Phys. Rev. B 23(12) (1981), pp. 6348–6356.
[133] H. Richter, Z.P. Wang, and L. Ley, Solid State Commun. 39(5) (1981), pp. 625–629.
[134] I.H. Campbell and P.M. Fauchet, Solid State Commun. 58(10) (1986), pp. 739–741.
[135] G.G. Samsonidze, R. Saito, A. Jorio, M.A. Pimenta, A.G. Souza Filho, A. Grüneis, G. Dresselhaus,
and M.S. Dresselhaus, J. Nanosci. Nanotechnol. 3(6) (2003), pp. 431–458.
[136] R. Saito, K. Sato, Y. Oyama, J. Jiang, G.G. Samsonidze, G. Dresselhaus, and M.S. Dresselhaus, Phys.
Rev. B 71 (2005), p. 153413.
[137] M.S. Dresselhaus, G. Dresselhaus, A.M. Rao, A. Jorio, A.G. Souza Filho, G.G. Samsonidze, and
R. Saito, Indian J. Phys. 77B (2003), pp. 75–99.
[138] R. Saito, J. Jiang, A. Grüneis, K. Sato, Y. Oyama, G.G. Samsonidze, S.G. Chou, G. Dresselhaus, M.S.
Dresselhaus, L.G. Cançado, C. Fantini, A. Jorio, and M.A. Pimenta, Mol. Cryst. Liquid Cryst. 455
(2006), pp. 287–294.
Advances in Physics 543
[139] J. Jiang, R. Saito, A. Grüneis, S.G. Chou, G.G. Samsonidze, A. Jorio, G. Dresselhaus, and
M.S. Dresselhaus, Phys. Rev. B 71 (2005), pp. 45417–45419.
[140] K. Nakada, M. Fujita, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 54 (1996), pp. 17954–17961.
[141] S. Reich, C. Thomsen, and J. Maultzsch, Carbon Nanotubes, Basic Concepts and Physical Properties,
Wiley-VCH, Berlin, 2004.
[142] E.B. Barros, A. Jorio, G.G. Samsonidze, R.B. Capaz, A.G. Souza Filho, J. Mendes Filho, G. Dressel-
haus, and M.S. Dresselhaus, Phys. Rep. 431(6) (2006), pp. 261–302.
[143] Y.W. Son, M.L. Cohen, and S.G. Louie, Phys. Rev. Lett. 97(21) (2006), p. 216803.
[144] Y.W. Son, M.L. Cohen, and S.G. Louie, Nature 444(7117) (2006), pp. 347–349.
[145] J. Maultzsch, S. Reich, and C. Thomsen, Raman scattering in carbon nanotubes revisited, Phys. Rev.
B 65 (2002), p. 3402.
[146] H. Telg, J. Maultzsch, S. Reich, and C. Thomsen, Phys. Rev. B 74(11) (2006), p. 115415.
[147] C.D. Spataru, S. Ismail-Beigi, L.X. Benedict, and S.G. Louie, Phys. Rev. Lett. 92(7) (2004), p. 77402.
[148] J. Jiang, R. Saito, G.G. Samsonidze, A. Jorio, S.G. Chou, G. Dresselhaus, and M.S. Dresselhaus, Phys.
Rev. B 75 (2007), pp. 35407–35413.
[149] V.G. Kravets, A.N. Grigorenko, R.R. Nair, P. Blake, S. Anissimova, K.S. Novoselov, and A.K. Geim,
Phys. Rev. B 81(15) (2010), p. 155413.
[150] M.Y. Sfeir, T. Beetz, F. Wang, L. Huang, X.M. Henry Huang, M. Huang, J. Hone, S. O’Brien, J.A.
Misewich, T.F. Heinz, L. Wu, Y. Zhu, and L.E. Brus, Science 312(5773) (2006), pp. 554–556.
[151] C. Casiraghi, A. Hartschuh, H. Qian, S. Piscanec, C. Georgi, A. Fasoli, K.S. Novoselov, D.M. Basko,
and A.C. Ferrari, Nano Lett. 9(4) (2009), pp. 1433–1441.
[152] M.J. O’Connell, S.M. Bachilo, C.B. Huffman, V.C. Moore, M.S. Strano, E.H. Haroz, K.L. Rialon,
P.J. Boul, W.H. Noon, and C. Kittrell, Science 297(5581) (2002), p. 593.
[153] S.M. Bachilo, L. Balzano, J.E. Herrera, F. Pompeo, D.E. Resasco, and R. Bruce Weisman, J. Am.
Chem. Soc. 125(37) (2003), pp. 11186–11187.
[154] S.G. Chou, F. Plentz, J. Jiang, R. Saito, D. Nezich, H.B. Ribeiro, A. Jorio, M.A. Pimenta,
G.G. Samsonidze, and A.P. Santos, Phys. Rev. Lett. 94(12) (2005), p. 127402.
[155] R.C.C. Leite and S.P.S. Porto, Phys. Rev. Lett. 17(1) (1966), pp. 10–12.
[156] M. Freitag, J. Chen, J. Tersoff, J.C. Tsang, Q. Fu, J. Liu, and Ph. Avouris, Phys. Rev. Lett. 93(7) (2004),
p. 76803.
[157] S. Essig, C.W. Marquardt, A. Vijayaraghavan, M. Ganzhorn, S. Dehm, F. Hennrich, F. Ou, A.A. Green,
C. Sciascia, F. Bonaccorso, K.-P. Bohnen, H.v. Lohneysen, M.M. Kappes, P.M. Ajayan, M.C. Hersam,
A.C. Ferrari, and R. Krupke, Nano Lett. 10 (2010), p. 1589.
[158] Y. Chen, R.C. Haddon, S. Fang, A.M. Rao, P.C. Eklund, W.H. Lee, E.C. Dickey, E.A. Grulke, J.C.
Pendergrass, and A. Chavan, J. Mater. Res. 13(9) (1998), pp. 2423–2431.
[159] U.J. Kim, C.A. Furtado, X. Liu, G. Chen, and P.C. Eklund, J. Am. Chem. Soc. 127(44) (2005),
pp. 15437–15445.
[160] J. Kastner, T. Pichler, H. Kuzmany, S. Curran, W. Blau, D.N. Weldon, M. Delamesiere, S. Draper, and
H. Zandbergen, Chem. Phys. Lett. 221(1–2) (1994), pp. 53–58.
[161] A. Gambetta, C. Manzoni, E. Menna, M. Meneghetti, G. Cerullo, G. Lanzani, S. Tretiak, A. Piryatinski,
A. Saxena, and R.L. Martin, Nat. Phys. 2(8) (2006), pp. 515–520.
[162] Y.-S. Lim, K.-J. Yee, J.-H. Kim, E.H. Hároz, J. Shaver, J. Kono, S.K. Doorn, R.H. Hauge, and R.E
Smalley, Nano Lett. 6(12) (2006), pp. 2696–2700.
[163] K. Ishioka, M. Hase, M. Kitajima, L. Wirtz, A. Rubio, and H. Petek, Phys. Rev. B 77 (2008),
p. 121402(R).
[164] K. Kato, K. Ishioka, M. Kitajima, J. Tang, R. Saito, and H. Petek, Nano Lett. 8(10) (2008), pp. 3102–
3108.
[165] A. Jones, A. Ballestad, T. Li, M. Whitwick, J. Rottler, and T. Tiedje, Phys. Rev. B 79(20) (2009),
pp. 205419–205434.
[166] K.S. Krishnan and C.V. Raman, Proc. R. Soc. Lond. Ser. A 115(772) (1927), pp. 549–554.
[167] M.S. Dresselhaus, G. Dresselhaus, R. Saito, and A. Jorio, Raman Spectroscopy of Carbon Nanotubes,
Vol. 409, Elsevier, the Netherlands, 2005.
[168] J. Klett, R. Hardy, E. Romine, C. Walls, and T. Burchell, Carbon 38(7) (2000), pp. 953–973.
544 R. Saito et al.
[169] A.C. Ferrari and J. Robertson, Phil. Trans. R. Soc. Lond. A 362 (2004), pp. 2477–2512.
[170] L.G. Cançado, M.A. Pimenta, B.R.A. Neves, M.S.S. Dantas, and A. Jorio, Phys. Rev. Lett. 93, 247401
(2004).
[171] A. Jorio, A.G. Souza Filho, G. Dresselhaus, M.S. Dresselhaus, A. Swan, M. Ünlü, B. Goldberg,
M.A. Pimenta, J. Hafner, C.M. Lieber, and R. Saito, Phys. Rev. B 65(15) (2002), p. 155412.
[172] A.C. Ferrari and J. Robertson, Phys. Rev. B 64(7) (2001), p. 75414.
[173] M.J. Pelletier, Analytical Applications of Raman Spectroscopy, Wiley-Blackwell, Oxford, UK, 1999,
p. 478.
[174] M.A. Pimenta, G. Dresselhaus, M.S. Dresselhaus, L.G. Cancado, A. Jorio, and R. Saito, Phys. Chem.
Chem. Phys. 9(11) (2007), pp. 1276–1290.
[175] C. Fantini, A. Jorio, M. Souza, M.S. Strano, M.S. Dresselhaus, and M.A. Pimenta, Phys. Rev. Lett.
93(14) (2004), p. 147406.
[176] Y. Kumazawa, H. Kataura, Y. Maniwa, I. Umezu, S. Suzuki, Y. Ohtsuka, and Y. Achiba, Synth. Met.
103(1–3) (1999), pp. 2555–2558.
[177] R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 61 (2000), pp. 2981–2990.
[178] A.M. Rao, P.C. Eklund, S. Bandow,A. Thess, and R.E. Smalley, Nature 388(6639) (1997), pp. 257–259.
[179] K. Kneipp, H. Kneipp, P. Corio, S. Brown, K. Shafer, J. Motz, L. Perelman, E. Hanlon, A. Marucci,
G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. Lett. 84(15) (2000), pp. 3470–3473.
[180] P. Corio, S.D.M. Brown, A. Marucci, M.A. Pimenta, K. Kneipp, G. Dresselhaus, and M.S. Dresselhaus,
Phys. Rev. B 61(19) (2000), pp. 13202–13211.
[181] S. Lefrant, Curr. Appl. Phys. 2(6) (2002), pp. 479–482.
[182] K. Kneipp, A. Jorio, H. Kneipp, S. Brown, K. Shafer, J. Motz, R. Saito, G. Dresselhaus, and
M.S. Dresselhaus, Phys. Rev. B 63(8) (2001), p. 81401.
[183] G. Goncalves, P.A.A.P. Marques, C.M. Granadeiro, H.I.S. Nogueira, M.K. Singh, and J. Gracio, Chem.
Mater. 21(20) (2009), pp. 4796–4802.
[184] L. Gao, W. Ren, B. Liu, R. Saito, Z.-S. Wu, S. Li, C. Jiang, F. Li, and H.-M. Cheng, ACS Nano 3(4)
(2009), pp. 933–9.
[185] X. Ling, L. Xie, Y. Fang, H. Xu, H. Zhang, J. Kong, M.S. Dresselhaus, J. Zhang, and Z. Liu, Nano
Lett. 10(2) (2010), pp. 553–561.
[186] F. Schedin, E. Lidorikis,A. Lombardo, V.G. Kravetsand,A.K. Geim,A.N. Grigorenko, K.S. Novoselov,
and A.C. Ferrari, ACS Nano 4 (2010), p. 5617.
[187] W. Ren, R. Saito, L. Gao, F. Zheng, Z. Wu, B. Liu, M. Furukawa, J. Zhao, Z. Chen, and H.-M. Cheng,
Phys. Rev. B 81(3) (2010), p. 35412.
[188] A. Hartschuh, H.N. Pedrosa, L. Novotny, and T.D. Krauss, Science 301(5638) (2003), p. 1354.
[189] K. Ikeda and K. Uosaki, Nano Lett. 9(4) (2009), pp. 1378–1381.
[190] A. Zumbusch, G.R. Holtom, and X. Sunney Xie, Phys. Rev. Lett. 82(20) (1999), pp. 4142–4145.
[191] A. Das, B. Chakraborty, S. Piscanec, S. Pisana, A.K. Sood, and A.C. Ferrari, Phys. Rev. B 79(15)
(2009), p. 155417.
[192] L.M. Malard, D.C. Elias, E.S. Alves, and M.A. Pimenta, Phys. Rev. Lett. 101(25) (2008),
p. 257401.
[193] S. Piscanec, M. Lazzeri, J. Robertson, A.C. Ferrari, and F. Mauri, Phys. Rev. B 75(3) (2007), p. 35427.
[194] S. Piscanec, M. Lazzeri, F. Mauri, A.C. Ferrari, and J. Robertson, Phys. Rev. Lett. 93(18) (2004),
p. 185503.
[195] A. Das, S. Pisana, B. Chakraborty, S. Piscanec, S.K. Saha, U.V. Waghmare, K.S. Novoselov, H.R.
Krishnamurthy, A.K. Geim, and A.C. Ferrari, Nature Nanotechnology, 3, 210–215, (2008).
[196] K. Sasaki, R. Saito, G. Dresselhaus, M.S. Dresselhaus, H. Farhat, and J. Kong, Phys. Rev. B 77 (2008),
p. 245441.
[197] M. Lazzeri and F. Mauri, Phys. Rev. Lett. 97(26) (2006), p. 266407.
[198] K. Ishikawa and T. Ando, J. Phys. Soc. Jpn. 75(8) (2006), p. 84713.
[199] V.N. Popov and P. Lambin, Phys. Rev. B 73(8) (2006), p. 854635607.
[200] J. Yan, E.A. Henriksen, P. Kim, and A. Pinczuk, Phys. Rev. Lett. 101(13) (2008), p. 136804.
[201] H. Farhat, H. Son, G. Samsonidze, S. Reich, M.S. Dresselhaus, and J. Kong, Phys. Rev. Lett. 99(14)
(2007), p. 145506.
Advances in Physics 545
[202] M. Kalbac, L. Kavan, H. Farhat, J. Kong, and M.S. Dresselhaus, J. Phys. Chem. C 113(5) (2009),
pp. 1751–1757.
[203] M. Kalbac, H. Farhat, L. Kavan, J. Kong, K. Sasaki, R. Saito, and M.S. Dresselhaus, ACS Nano 3
(2009), pp. 2320–2328.
[204] H. Farhat, K. Sasaki, M. Kalbac, M. Hofmann, R. Saito, M.S. Dresselhaus, and J. Kong, Phys. Rev.
Lett. 102(12) (2009), pp. 268041–4.
[205] K. Sasaki, H. Farhat, R. Saito, and M.S. Dresselhaus, Phys. E 42(8) (2010), pp. 2005–2015.
[206] H. Farhat, S. Berciaud, M. Kalbac, R. Saito, T. Heinz, M.S. Dresselhaus, and J. Kong, unpublished.
[207] A.V. Baranov, A.N. Bekhterev,Y.S. Bobovich, and V.I. Petrov, Opt. Spectrosc. 62 (1987), pp. 612–616.
[208] J. Maultzsch, S. Reich, and C. Thomsen, Phys. Rev. B 70(15) (2004), p. 155403.
[209] R. Saito, A. Grüneis, G.G. Samsonidze, V.W. Brar, G. Dresselhaus, M.S. Dresselhaus, A. Jorio, L.G.
Cançado, C. Fantini, and M.A. Pimenta, New J. Phys. 5 (2003), p. 157.
[210] R. Saito, A. Jorio, A.G. Souza Filho, G. Dresselhaus, M.S. Dresselhaus, and M. Pimenta, Phys. Rev.
Lett. 88(2) (2001), p. 27401.
[211] R. Saito, A. Grüneis, L.G. Cançado, M.A. Pimenta, A. Jorio, A.G. Souza Filho, M.S. Dresselhaus, and
G. Dresselhaus, Mol. Cryst. Liq. Cryst. 387 (2002), pp. 287–296.
[212] I. Pócsik, M. Hundhausen, M. Koósa, and L. Ley, J. Non-Cryst. Solids 227–230 (1998), pp. 1083–1086.
[213] D. Mafra, G. Samsonidze, L. Malard, D. Elias, J. Brant, F. Plentz, E. Alves, and M. Pimenta, Phys.
Rev. B 76(23) (2007) 233407.
[214] R. Vidano and D.B. Fischbach, J. Am. Ceram. Soc. 61(1–2) (1978), pp. 13–17.
[215] R.P. Vidano and L.J. Fischbach, Solid State Commun. 39(2) (1981), pp. 341–344.
[216] L.G. Cançado, M.A. Pimenta, R. Saito, A. Jorio, L.O. Ladeira, A. Grueneis, A.G. Souza-Filho, G.
Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 66(3) (2002), p. 35415.
[217] J. Zimmermann, P. Pavone, and G. Cuniberti, Phys. Rev. B 78(4) (2008), p. 45410.
[218] M. Lazzeri, C. Attaccalite, L. Wirtz, and F. Mauri, Phys. Rev. B 78(8) (2008), p. 81406.
[219] D.M. Basko, Phys. Rev. B 78(11) (2008), p. 115432.
[220] D.M. Basko, S. Piscanec, and A.C. Ferrari, Phys. Rev. B 80(16) (2009), p. 165413.
[221] T. Shimada, T. Sugai, C. Fantini, M. Souza, L.G. Cançado,A. Jorio, M.A. Pimenta, R. Saito,A. Grüneis,
and G. Dresselhaus, Carbon 43(5) (2005), pp. 1049–1054.
[222] M.A. Pimenta, E.B. Hanlon, A. Marucci, P. Corio, S.D.M. Brown, S.A. Empedocles, M.G. Bawendi,
G. Dresselhaus, and M.S. Dresselhaus, Brazilian J. Phys. 30(2) (2000), pp. 423–427.
[223] R.J. Nemanich and S.A. Solin, Phys. Rev. B 20(2) (1979), pp. 392–401.
[224] P.H. Tan, Y.M. Deng, and Q. Zhao, Phys. Rev. B 58(9) (1998), pp. 5435–5439.
[225] E.F. Atunes, A.O. Lobo, E.J. Corat, V.J. Trava-Airoldi, A.A. Martin, and C. Verissimo, Carbon 44(11)
(2006), pp. 2202–2211.
[226] P.H. Tan, C.Y. Hu, J. Dong, W.C. Shen, and B.F. Zhang, Phys. Rev. B 64(21) (2001), p. 214301.
[227] V.W. Brar, G.G. Samsonidze, M.S. Dresselhaus, G. Dresselhaus, R. Saito, A. Swan, M. Ünlü,
B. Goldberg, A.G. Souza Filho, and A. Jorio, Phys. Rev. B 66(15) (2002), p. 155418.
[228] M. Pimenta, A. Marucci, S. Empedocles, M. Bawendi, E. Hanlon, A. Rao, P. Eklund, R. Smalley,
G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 58(24) (1998), pp. R16016–R16019.
[229] P.T. Araujo, A. Jorio, M.S. Dresselhaus, K. Sato, and R. Saito, Phys. Rev. Lett. 103(14) (2009),
p. 146802.
[230] P.T. Araujo, P.B.C. Pesce, M.S. Dresselhaus, K. Sato, R. Saito, and A. Jorio, Phys. E 42 (2010), pp.
1251–1261.
[231] Y. Miyauchi, R. Saito, K. Sato, Y. Ohno, S. Iwasaki, T. Mizutani, J. Jiang, and S. Maruyama, Chem.
Phys. Lett. 442(4–6) (2007), pp. 394–399.
[232] A.R.T. Nugraha, R. Saito, K. Sato, P.T. Araujo, A. Jorio, and M.S. Dresselhaus, Appl. Phys. Lett. 97(9)
(2010), p. 91905.
[233] K. Sato, R. Saito, Y. Oyama, J. Jing, L.G. Cancado, M.A. Pimenta, A. Jorio, G.G. Samsonidze, G.
Dresselhaus, and M.S. Dresselhaus, Chem. Phys. Lett. 427(1–3) (2006), pp. 117–121.
[234] S. Uryu and T. Ando, Phys. Semicond. B 893 (2007), pp. 1033–1034.
[235] R.A. Jishi, L. Venkataraman, M.S. Dresselhaus, and G. Dresselhaus, Chem. Phys. Lett. 209 (1993),
pp. 77–82.
546 R. Saito et al.
[236] R. Nicklow, N. Wakabayashi, and H.G. Smith, Phys. Rev. B 5(12) (1972), pp. 4951–4962.
[237] J.L. Wilkes, R.E. Palmer, and R.F. Willis, J. Electron Spectrosc. Rel. Phenom. 44(1) (1987), pp. 355–
360.
[238] T. Aizawa, R. Souda, S. Otani, Y. Ishizawa, and C. Oshima, Phys. Rev. B 42(18) (1990), pp. 11469–
11478.
[239] R. Al-Jishi and G. Dresselhaus, Phys. Rev. B 26(8) (1982), pp. 4514–4522.
[240] S. Siebentritt, R. Pues, K.-H. Rieder, and A.M. Shikin, Phys. Rev. B 55(12) (1997), pp. 7927–7934.
[241] C. Mapelli, C. Castiglioni, G. Zerbi, and K. Müllen, Phys. Rev. B 60(18) (1999), pp. 12710–
12725.
[242] C.D. Spataru, S. Ismail-Beigi, L.X. Benedict, and S.G. Louie, Appl. Phys. A Mater. Sci. Process. 78(8)
(2004), pp. 1129–1136.
[243] R. Saito, A. Jorio, J. Jiang, K. Sasaki, G. Dresselhaus, and M.S. Dresselhaus, Optical Properties of
Carbon Nanotubes and Nanographene, Oxford University Press, Oxford, UK, 2010.
[244] A. Grüneis, J. Serrano, A. Bosak, M. Lazzeri, S.L. Molodtsov, L. Wirtz, C. Attaccalite, M. Krisch, A.
Rubio, F. Mauri, and T. Pichler, Phys. Rev. B 80(8) (2009), p. 85423.
[245] M. Furukawa, Thesis. Master’s Thesis, Tohoku University, 2010.
[246] K. Sasaki, M. Yamamoto, S. Murakami, R. Saito, M.S. Dresselhaus, K. Takai, T. Mori, T. Enoki, and
K. Wakabayashi, Phys. Rev. B 80(15) 155450 (2009).
[247] N. Mounet and N. Marzari, Phys. Rev. B 71(20) (2005), p. 205214.
[248] D. Sanchez-Portal, E.Artacho, J.M. Solar,A. Rubio, and P. Ordejon, Phys. Rev. B 59 (1999), pp. 12678–
12688.
[249] O. Dubay and G. Kresse, Phys. Rev. B 67(3) (2003), p. 35401.
[250] J.-C. Charlier, P. Eklund, J. Zhu, and A.C. Ferrari, Electron and phonon properties of graphene: their
relationship with carbon nanotubes, Vol. 111, Springer Series on Topics in Applied Physics, Springer,
Berlin, 2008, pp. 625–680.
[251] P.H. Tan, L. An, L.Q. Liu, Z.X. Guo, R. Czerw, D.L. Carroll, P.M. Ajayan, N. Zhang, and H.L. Guo,
Phys. Rev. B 66(24) (2002), p. 245410.
[252] J. Hone, Phonons and Thermal Properties of Carbon Nanotubes, Vol. 80, Springer, Berlin, 2001,
pp. 273–286.
[253] D. Porezag, Th. Frauenheim, Th. Köhler, G. Seifert, and R. Kaschner, Phys. Rev. B 51(19) (1995),
pp. 12947–12957.
[254] G.G. Samsonidze, R. Saito, N. Kobayashi, A. Grüneis, J. Jiang, A. Jorio, S.G. Chou, G. Dresselhaus,
and M.S. Dresselhaus, Appl. Phys. Lett. 85 (2004), pp. 5703–5705.
[255] A. Grüneis, C. Attaccalite, L. Wirtz, H. Shiozawa, R. Saito, T. Pichler, and A. Rubio, Phys. Rev. B
78(20) (2008), p. 205425.
[256] S. Reich, J. Maultzsch, C. Thomsen, and P. Ordejón, Phys. Rev. B 66(3) (2002), p. 35412.
[257] J. Slonczewski and P. Weiss, Phys. Rev. 109(2) (1958), pp. 272–279.
[258] M. Koshino and T. Ando, Phys. Rev. B 77(11) (2008), p. 115313.
[259] J.C. Slater and G.F. Koster, Phys. Rev. 94(6) (1954), pp. 1498–1524.
[260] J.J.P. Stewart, MOPAC93.00, Fujitsu Limited, Tokyo, Japan, 1993.
[261] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V.
Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov,
J. Bloino, G. Zheng, and D.J. Sonnenb, Gaussian 09, Revision A.1, Gaussian, Inc., Wallingford, CT,
2009.
[262] A. Grüneis, R. Saito, G.G. Samsonidze, T. Kimura, M.A. Pimenta, A. Jorio, A.G. Souza Filho,
G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 67(16) (2003), pp. 165402–165407.
[263] Y. Oyama, R. Saito, K. Sato, J. Jiang, G.G. Samsonidze, A. Grüneis, Y. Miyauchi, S. Maruyama, A.
Jorio, G. Dresselhaus, and M.S. Dresselhaus, Carbon 44(5) (2006), pp. 873–879.
[264] R. Saito, A. Grüneis, G.G. Samsonidze, G. Dresselhaus, M.S. Dresselhaus, A. Jorio, L.G. Cancado,
M.A. Pimenta, and A.G. Souza Filho, Appl. Phys. A 78(8) (2004), pp. 1099–1105.
[265] J. Jiang, R. Saito,A. Grüneis, G. Dresselhaus, and M.S. Dresselhaus, Carbon 42 (2004), pp. 3169–3176.
[266] G.G. Samsonidze, E.B. Barros, R. Saito, J. Jiang, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev.
B 75 (2007), p. 155420.
Advances in Physics 547
[267] J. Jiang, R. Saito, A. Grüneis, G. Dresselhaus, and M.S. Dresselhaus, Chem. Phys. Lett. 392 (2004),
pp. 383–389.
[268] J. Jiang, R. Saito, G.G. Samsonidze, S.G. Chou, A. Jorio, G. Dresselhaus, and M.S. Dresselhaus, Phys.
Rev. B 72 (2005), pp. 235408–235411.
[269] J. Jiang, R. Saito, A. Grüneis, S. Chou, G. Samsonidze, A. Jorio, G. Dresselhaus, and M.S. Dresselhaus,
Phys. Rev. B 71(20) (2005), p. 205420.
[270] A. Grüneis, Resonance Raman spectroscopy of single wall carbon nanotubes, Ph.D. thesis, Tohoku
University, Sendai, Japan, September 2004.
[271] R. Saito and H. Kamimura, J. Phys. Soc. Jpn. 52(2) (1983), p. 407.
[272] T. Ando, J. Phys. Soc. Jpn. 74(3) (2005), pp. 777–817.
[273] J. Jiang, R. Saito, K. Sato, J. Park, G.G. Samsonidze, A. Jorio, G. Dresselhaus, and M.S. Dresselhaus,
Phys. Rev. B 75(3) (2007), p. 35405.
[274] M. Rohlfing and S.G. Louie, Phys. Rev. B 62(8) (2000), pp. 4927–4944.
[275] V. Perebeinos, J. Tersoff, and P. Avouris, Phys. Rev. Lett. 92(25) (2004), p. 257402.
[276] R.B. Capaz, C. Spataru, S. Ismail-Beigi, and S.G. Louie, Phys. Rev. B 74(12) (2006), pp. 5–8.
[277] E.B. Barros, R.B. Capaz, A. Jorio, G.G. Samsonidze, A.G. Souza Filho, S. Ismail-Beigi, C.D. Spataru,
S.G. Louie, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 73(24) (2006), p. 241406.
[278] J.S. Park, Y. Oyama, R. Saito, W. Izumida, J. Jiang, K. Sato, C. Fantini, A. Jorio, G. Dresselhaus, and
M.S. Dresselhaus, Phys. Rev. B 74 (2006), p. 165414.
[279] Z. Yu and L.E. Brus, J. Phys. Chem. B 105(29) (2001), pp. 6831–6837.
[280] S. Reich, C. Thomsen, and P. Ordejón, Phy. Rev. B, 64, (2001) p. 195416.
[281] P.K. Eklund, G. Dresselhaus, M.S. Dresselhaus, and J. Fischer, Phys. Rev. B 16(8) (1977), pp. 3330–
3333.
[282] A.W. Bushmaker, V.V. Deshpande, S. Hsieh, M.W. Bockrath, and S.B. Cronin, Nano Lett. 9(8) (2009),
pp. 2862–2866.
[283] Y. Wu, J. Maultzsch, E. Knoesel, B. Chandra, M. Huang, M. Sfeir, L.E. Brus, J. Hone, and T.F. Heinz,
Phys. Rev. Lett. 99(2) (2007), p. 27402.
[284] L.M. Malard, J. Nilsson, D.C. Elias, J.C. Brant, F. Plentz, E.S. Alves, A.H. Castro-Neto, and M.A.
Pimenta, Phys. Rev. B 76(20) (2007), p. 201401.
[285] N. Ferralis, R. Maboudian, and C. Carraro, Phys. Rev. Lett. 101(15) (2008), p. 156801.
[286] T.M.G. Mohiuddin, A. Lombardo, R.R. Nair, A. Bonetti, G. Savini, R. Jalil, N. Bonini, D.M. Basko, C.
Galiotis, N. Marzari, K.S. Novoselov, A.K. Geim, and A.C. Ferrari, Phys. Rev. B 79 (2009), p. 205433.
[287] M. Huang, H. Yan, C. Chen, D. Song, T.F. Heinz, and J. Hone, Phonon softening and crystallographic
orientation of strained graphene studied by Raman spectroscopy, May 2009.
[288] S. Pisana, M. Lazzeri, C. Casiraghi, K.S. Novoselov, A.K. Geim, A.C. Ferrari, and F. Mauri, Nat.
Mater. 6(3) (2007), pp. 198–201.
[289] J. Yan, Y. Zhang, P. Kim, and A. Pinczuk, Phys. Rev. Lett. 98(16) 166802 (2007).
[290] Y. Wang, Z. Ni, T. Yu, Z.X. Shen, H. Wang, Y. Wu, W. Chen, and A.T. Shen Wee, J. Phys. Chem. C
112(29) (2008), pp. 10637–10640.
[291] C. Casiraghi, S. Pisana, K.S. Novoselov, A.K. Geim, and A.C. Ferrari, Appl. Phys. Lett. 91 (2007),
p. 233108.
[292] G. Compagnini, F. Giannazzo, S. Sonde, V. Raineri, and E. Rimini, Carbon 47(14) (2009), pp. 3201–
3207.
[293] I.O. Maciel, N. Anderson, M.A. Pimenta, A. Hartschuh, H. Qian, M. Terrones, H. Terrones, J. Campos-
Delgado, A.M. Rao, L. Novotny, and A. Jorio, Nat. Mater. 7(11) (2008), pp. 878–883.
[294] X. Wang, L. Zhi, and K. Müllen, Nano Lett. 8(1) (2008), pp. 323–327.
[295] Z. Ni, Y. Wang, T. Yu, Y. You, and Z. Shen, Phys. Rev. B 77(23) (2008), p. 235403.
[296] D.M. Basko, Phys. Rev. B 76(8) (2007), p. 81405.
[297] J.S. Park, A. Reina Cecco, R. Saito, J. Jiang, G. Dresselhaus, and M.S. Dresselhaus, Carbon 47 (2009),
pp. 1303–1310.
[298] L.G. Cançado, A. Reina, J. Kong, and M.S. Dresselhaus, Phys. Rev. B 77(24) (2008), p. 245408.
[299] P. Lespade, A. Marchand, M. Couzi, and F. Cruege, Carbon 22(4–5) (1984), pp. 375–385.
[300] P. Lespade, R. Al-Jishi, and M.S. Dresselhaus, Carbon 20(5) (1982), pp. 427–431.
[301] H. Wilhelm, M. Lelaurain, E. McRae, and B. Humbert, J. Appl. Phys. 84 (1998), p. 6552.
548 R. Saito et al.
[302] R.J. Nemanich and S.A. Lucovsky, Solid State Commun. 23(2) (1977), pp. 117–120.
[303] L.G. Cancado, K. Takai, T. Enoki, M. Endo, Y.A. Kim, H. Mizusaki, A. Jorio, L.N. Coelho,
R. Magalhaes-Paniago, and M.A. Pimenta, Appl. Phys. Lett. 88(16) (2006), p. 163106.
[304] P. Poncharal, A. Ayari, T. Michel, and J.L. Sauvajol, Phys. Rev. B 79(19) (2009), p. 195417.
[305] A. Reina, H.B. Son, L.Y. Jiao, B. Fan, M.S. Dresselhaus, Z.F. Liu, and J. Kong, J. Phys. Chem. C
112(46) (2008), pp. 17741–17744.
[306] A. Reina, S. Thiele, X. Jia, S. Bhaviripudi, M.S. Dresselhaus, J.A. Schaefer, and J. Kong, Nano Res.
2 (2009), pp. 509–516.
[307] C.H. Lui, Z. Li, Z. Chen, P.V. Klimov, L.E. Brus, and T.F. Heinz, Nano Lett. 11(1) (2011),
pp. 164–169.
[308] M.S. Dresselhaus and R. Kalish, Ion Implantation in Diamond, Graphite and Related Materials,
Springer Series in Materials Science, Springer, Berlin, 1992.
[309] A. Jorio, M.M. Lucchese, F. Stavale, E.H. Martins Ferreira, M.V.O. Moutinho, R.B. Capaz, and
C.A. Achete, J. Phys.: Condens. Matter 22(33) (2010), p. 334204.
[310] A. Grüneis, R. Saito, T. Kimura, L.G. Cançado, M.A. Pimenta, A. Jorio, A.G. Souza Filho,
G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 65(15) (2002), pp. 155405–155407.
[311] A.K. Gupta, T.J. Russin, H.R. Gutierrez, and P.C. Eklund, ACS Nano 3 (2009), pp. 45–52.
[312] B. Krauss, P. Nemes-Incze, V. Skakalova, L.P. Biro, K. von Klitzing, and J.H. Smet, Nano Lett. (11)(10)
(2010), pp. 4544–4548.
[313] M.S. Dresselhaus, G. Dresselhaus, and A. Jorio, J. Phys. Chem. C 111 (2007), pp. 17887–17893.
[314] P.T. Araujo, I.O. Maciel, P.B.C. Pesce, M.A. Pimenta, S.K. Doorn, H. Qian, A. Hartschuh, M. Steiner,
L. Grigorian, and K. Hata, Phys. Rev. B 77(24) (2008), p. 241403.
[315] G.D. Mahan, Phys. Rev. B 65(23) (2002), p. 235402.
[316] K. Hata, D.N. Futaba, K. Mizuno, T. Namai, M. Yumura, and S. Iijima, Science 306(5700) (2004),
p. 1362.
[317] S.M. Bachilo, M.S. Strano, C. Kittrell, R.H. Hauge, R.E. Smalley, and R.B. Weisman, Science
298(5602) (2002), p. 2361.
[318] M.S. Strano, J. Am. Chem. Soc. 125(51) (2003), pp. 16148–16153.
[319] S.K. Doorn, D.A. Heller, P.W. Barone, M.L. Usrey, and M.S. Strano, Appl. Phys. A 78(8) (2004),
pp. 1147–1155.
[320] T. Michel, M. Paillet, P. Poncharal, A. Zahab, J.-L. Sauvajol, J.C. Meyer, and S. Roth, Carbon
Nanotubes 222(II) (2006), pp. 121–122.
[321] P.T. Araujo, S.K. Doorn, S. Kilina, S. Tretiak, E. Einarsson, S. Maruyama, H. Chacham, M.A. Pimenta,
and A. Jorio, Phys. Rev. Lett. 98(6) (2007), p. 67401.
[322] P.T. Araujo, C. Fantini, M.M. Lucchese, M.S. Dresselhaus, and A. Jorio, Appl. Phys. Lett. 95(26)
(2009), p. 261902.
[323] J.S. Soares, L.G. Cançado, E.B. Barros, and A. Jorio, Phys. Stat. Solidi B 247(11–12) (2010), pp. 2835–
2837.
[324] J.S. Soares, A.P.M. Barboza, P.T. Araujo, N.M. Barbosa Neto, D. Nakabayashi, N. Shadmi, T.S.Yarden,
A. Ismach, N. Geblinger, E. Joselevich, C. Vilani, L.G. Cancado, L. Novotny, G. Dresselhaus, M.S.
Dresselhaus, B.R.A. Neves, M.S.C. Mazzoni, and A. Jorio, Nano Lett. 10(12) (2010), pp. 5043–5048.
[325] R. Pfeiffer, C. Kramberger, F. Simon, H. Kuzmany, V.N. Popov, and H. Kataura, Eur. Phys. J. B 42(3)
(2004), pp. 345–350.
[326] R. Pfeiffer, F. Simon, H. Kuzmany, and V.N. Popov, Phys. Rev. B 72(16) (2005), p. 161404.
[327] R. Pfeiffer, F. Simon, H. Kuzmany, V.N. Popov, V. Zolyomi, and J. Kurti, Phys. Stat. Solidi B 243(13)
(2006), pp. 3268–3272.
[328] R. Pfeiffer, H. Peterlik, H. Kuzmany, F. Simon, K. Pressl, P. Knoll, M.H. Rümmeli, H. Shiozawa, H.
Muramatsu, Y.A. Kim, T. Hayashi, and M. Endo, Phys. Stat. Solidi B 245(10) (2008), pp. 1943–1946.
[329] H. Kuzmany, W. Plank, R. Pfeiffer, and F. Simon, J. Raman Spectrosc. 39(2) (2008), pp. 134–140.
[330] F. Villalpando-Paez, H. Muramatsu,Y.A. Kim, H. Farhat, M. Endo, M. Terrones, and M.S. Dresselhaus,
Nanoscale 2(3) (2010), pp. 406–411.
[331] F. Villalpando-Paez, H. Son, D. Nezich, Y.P. Hsieh, J. Kong, Y.A. Kim, D. Shimamoto, H. Muramatsu,
T. Hayashi, M. Endo, M. Terrones, and M.S. Dresselhaus, Nano Lett. 8 (2008), pp. 3879–3886.
Advances in Physics 549
[332] X. Zhao, Y. Ando, L.-C. Qin, H. Kataura, Y. Maniwa, and R. Saito, Chem. Phys. Lett. 361(1–2) (2002),
pp. 169–174.
[333] M.S. Dresselhaus and P.C. Eklund, Adv. Phys. 49 (2000), pp. 705–814.
[334] S.D.M. Brown, P. Corio, A. Marucci, M.S. Dresselhaus, M.A. Pimenta, and K. Kneipp, Phys. Rev. B
61(8) (2000), pp. 5137–5140.
[335] P.T. Araujo and A. Jorio, Phys. Stat. Solidi B 245(10) (2008), pp. 2201–2204.
[336] G.S. Duesberg, W.J. Blau, H.J. Byrne, J. Muster, M. Burghard, and S. Roth, Chem. Phys. Lett. 310(1–2)
(1999), pp. 8–14.
[337] J. Azoulay, A. Débarre, A. Richard, and P. Tchénio, J. Phys. IV 10(8) (2000), pp. 8–223.
[338] A. Jorio, A.G. Souza Filho, G. Dresselhaus, M.S. Dresselhaus, R. Saito, J. Hafner, C. Lieber,
F. Matinaga, M. Dantas, and M.A. Pimenta, Phys. Rev. B 63(24) (2001), p. 245416.
[339] J.H. Hafner, C.L. Cheung, T.H. Oosterkamp, and C.M. Lieber, J. Phys. Chem. B 105(4) (2001),
pp. 743–746.
[340] A. Jorio, C. Fantini, M.S.S. Dantas, M.A. Pimenta, A.G. Souza Filho, G.G. Samsonidze, V.W. Brar,
G. Dresselhaus, M.S. Dresselhaus, A. Swan, M. Ünlü, B. Goldberg, and R. Saito, Phys. Rev. B 66(11)
(2002), p. 115411.
[341] A.G. Souza Filho, A. Jorio, J. Hafner, C. Lieber, R. Saito, M. Pimenta, G. Dresselhaus, and M.S.
Dresselhaus, Phys. Rev. B 63(24) (2001), p. 241404R.
[342] K. Sato, R. Saito, J. Jiang, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 76 (2007), p. 195446.
[343] K. Sato, R. Saito, J. Jiang, G. Dresselhaus, and M.S. Dresselhaus, Vib. Spectrosc. 45(2) (2007),
pp. 89–94.
[344] P.B.C. Pesce, P.T. Araujo, P. Nikolaev, S.K. Doorn, K. Hata, R. Saito, M.S. Dresselhaus, and A. Jorio,
Appl. Phys. Lett. 96 (2010), p. 51910.
[345] S. Doorn, P. Araujo, K. Hata, and A. Jorio, Phys. Rev. B, 78, 165408. (2008).
[346] V.N. Popov, New J. Phys. 6 (2004), p. 17.
[347] S. Maruyama, R. Kojima, Y. Miyauchi, S. Chiashi, and M. Kohno, Chem. Phys. Lett. 360(3–4) (2002),
pp. 229–234.
[348] P. Nikolaev, M.J. Bronikowski, R.K. Bradley, F. Rohmund, D.T. Colbert, K.A. Smith, and R.E. Smalley,
Chem. Phys. Lett. 313(1–2) (1999), pp. 91–97.
[349] T. Ando, J. Phys. Soc. Jpn. 79(2) (2010), p. 4706.
[350] M.S. Dresselhaus, G. Dresselhaus, A. Jorio, A.G. Souza Filho, and R. Saito, Carbon 40(12) (2002),
pp. 2043–2061.
[351] A. Jorio, G. Dresselhaus, M.S. Dresselhaus, M. Souza, M.S.S. Dantas, M.A. Pimenta, A.M. Rao, R.
Saito, C. Liu, and H.M. Cheng, Phys. Rev. Lett. 85(12) (2000), pp. 2617–2620.
[352] A. Jorio, M.A. Pimenta, A.G. Souza Filho, G.G. Samsonidze, A.K. Swan, M.S. Ünlü, B.B. Goldberg,
R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. Lett. 90(10) (2003), p. 107403.
[353] H. Ajiki and T. Ando, Phys. B: Condens. Matter 201 (1994), pp. 349–352.
[354] A.G. Marinopoulos, L. Reining, A. Rubio, and N. Vast, Phys. Rev. Lett. 91(4) (2003), p. 46402.
[355] G.S. Duesberg, I. Loa, M. Burghard, K. Syassen, and S. Roth, Phys. Rev. Lett. 85(25) (2000),
pp. 5436–5439.
[356] J. Hwang, H.H. Gommans, A. Ugawa, and H. Tashiro, Phys. Rev. B 62 (2000), pp. R13310–R13313.
[357] A. Jorio, A.G. Souza Filho, V. Brar, A. Swan, M. Ünlü, B. Goldberg, A. Righi, J. Hafner, C.M. Lieber,
R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 65(12) (2002), p. R121402.
[358] A. Rao, A. Jorio, M.A. Pimenta, M. Dantas, R. Saito, G. Dresselhaus, and M.S. Dresselhaus, Phys.
Rev. Lett. 84(8) (2000), pp. 1820–1823.
[359] C. Thomsen, S. Reich, P.M. Rafailov, and H. Jantoliak, Phys. Stat. Solidi B 214 (1999),
pp. 15–16.
[360] M. Souza, A. Jorio, C. Fantini, B.R.A. Neves, M.A. Pimenta, R. Saito, A. Ismach, E. Joselevich, V.W.
Brar, G.G. Samsonidze, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 69(24) (2004), p. R15424
1–4.
[361] P. Corio, A. Jorio, N. Dimer, and M.S. Dresselhaus, Chem. Phys. Lett 392 (2004), pp. 396–402.
[362] L. Kavan, L. Dunsch, H. Kataura, A. Oshiyama, M. Otani, and S. Okada, J. Phys. Chem. B 107(31)
(2003), pp. 7666–7675.
550 R. Saito et al.
[363] A.W. Bushmaker, V.V. Deshpande, S. Hsieh, M.W. Bockrath, and S.B. Cronin, Nano Lett. 9(2) (2009),
pp. 607–611.
[364] R. Saito, A. Jorio, J. Hafner, C.M. Lieber, M. Hunter, T. McClure, G. Dresselhaus, and M.S. Dressel-
haus, Phys. Rev. B 64(8) (2001), pp. 85312–85319.
[365] K. Sasaki, R. Saito, G. Dresselhaus, M.S. Dresselhaus, H. Farhat, and J. Kong, Phys. Rev. B 78 (2008),
pp. 235405–235411.
[366] J.C. Tsang, M. Freitag, V. Perebeinos, J. Liu, and P. Avouris, Nat. Nanotechnol. 2(11) (2007),
pp. 725–730.
[367] A.G. Souza-Filho, A. Jorio, G.G. Samsonidze, G. Dresselhaus, R. Saito, and M.S. Dresselhaus,
Nanotechnology 14 (2003), pp. 1130–1139.
[368] N. Geblinger, A. Ismach, and E. Joselevich, Nat. Nanotechnol. 3(4) (2008), pp. 195–200.
[369] A. Jorio, C. Fantini, M.A. Pimenta, R.B. Capaz, G.G. Samsonidze, G. Dresselhaus, M.S. Dresselhaus,
J. Jiang, N. Kobayashi, and A. Grüneis, Phys. Rev. B 71(7) (2005), p. 75401.
[370] J. Kürti, V. Zólyomi, M. Kertesz, and G.Y. Sun, New J. Phys. 5 (2003), p. 125.
[371] C. Fantini, A. Jorio, M. Souza, and L.O. Ladeira, Phys. Rev. Lett. 93 (2004), p. 87401.
[372] C. Fantini, A. Jorio, M. Souza, R. Saito, G.G. Samsonidze, M.S. Dresselhaus, and M.A. Pimenta1, in
Proceedings of the XVIII International Winter School on the Electronic Properties of Novel Materials,
H. Kuzmany, J. Fink, M. Mehring, and S. Roth, eds., Vol. 786, pp. 178–181, American Institute of
Physics, Woodbury, NY, 2005.
[373] R. Saito, T. Takeya, T. Kimura, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 59(3) (1999),
pp. 2388–2392.
[374] S.G. Chou, H. Son, M. Zheng, R. Saito, A. Jorio, G. Dresselhaus, and M.S. Dresselhaus, Chem. Phys.
Lett. 443(4–6) (2007), pp. 328–332.
[375] A.G. Souza Filho, A. Jorio, A.K. Swan, M.S. Ünlü, B.B. Goldberg, R. Saito, J. Hafner, C.M. Lieber,
M.A. Pimenta, G. Dresselhaus, and M.S. Dresselhaus, Phys. Rev. B 65(8) (2002), p. 85417.
[376] A.G. Souza Filho, A. Jorio, G. Dresselhaus, M.S. Dresselhaus, R. Saito, A.K. Swan, M.S. Ünlü,
B.B. Goldberg, J.H. Hafner, C.M. Lieber, and M.A. Pimenta, Phys. Rev. B 65(3) (2001),
p. 35404.
[377] A.G. Souza-Filho, A. Jorio, G.G. Samsonidze, G. Dresselhaus, M.A. Pimenta, M.S. Dresselhaus, A.K.
Swan, M.S. Ünlü, B.B. Goldberg, and R. Saito, Phys. Rev. B 67 (2003), p. 035427-1–7.
[378] G.G. Samsonidze, R. Saito, A. Jorio, A.G. Souza-Filho, A. Grüneis, M.A. Pimenta, G. Dresselhaus,
and M.S. Dresselhaus, Phys. Rev. Lett. 90 (2003), p. 27403.
[379] A.G. Souza Filho, A. Jorio, G.G. Samsonidze, G. Dresselhaus, M.S. Dresselhaus, A.K. Swan, M. Ünlü,
B.B. Goldberg, R. Saito, and J.H. Hafner, Chem. Phys. Lett. 354(1–2) (2002), pp. 62–68.
[380] R. Saito, A. Jorio, A.G. Souza Filho, G. Dresselhaus, M.S. Dresselhaus, A. Grüneis, L.G. Cançado,
and M.A. Pimenta, Jpn. J. Appl. Phys. 41(Part 1, No. 7B) (2002), pp. 4878–4882.
[381] L. Novotny and B. Hecht, Principles of Nanooptics, Cambridge University Press, 2006, p. 558.
[382] A. Hartschuh, E.J. Sánchez, X.S. Xie, and L. Novotny, Phys. Rev. Lett. 90(9) (2003), p. 95503.
[383] I.O. Maciel, M.A. Pimenta, M. Terrones, H. Terrones, J. Campos-Delgado, and A. Jorio, Phys. Stat.
Solidi B 245(10) (2008), pp. 2197–2200.
[384] I.O. Maciel, J. Campos-Delgado, E. Cruz-Silva, M.A. Pimenta, B.G. Sumpter, V. Meunier, F. López-
Urias, E. Munoz-Sandoval, H. Terrones, M. Terrones, and A. Jorio Nano Lett. 9(6) (2009), pp. 2267–
2272.
[385] I.O. Maciel, J. Campos-Delgado, M.A. Pimenta, M. Terrones, H. Terrones, A.M. Rao, and A. Jorio,
Phys. Stat. Solidi B 246(11–12) (2009), pp. 2432–2435.
[386] C. Georgi and A. Hartschuh, Appl. Phys. Lett. 97(14) (2010), p. 143117.
[387] H. Qian, C. Georgi, N. Anderson, A.A. Green, M.C. Hersam, L. Novotny, and A. Hartschuh, Nano
Lett. 8(5) (2008), pp. 1363–1367.
[388] N. Anderson, A. Hartschuh, and L. Novotny, Nano Lett. 7(3) (2007), pp. 577–582.
[389] L.G. Cançado, A. Hartschuh, and L. Novotny, J. Raman Spectrosc. 40(10) (2009), pp. 1420–1426.
[390] C.H. Park, L. Yang, Y.W. Son, M.L. Cohen, and S.G. Louie, Nat. Phys. 4(3) (2008), pp. 213–217.
[391] A. Jorio, C. Fantini, M.A. Pimenta, D.A. Heller, M.S. Strano, M.S. Dresselhaus, Y. Oyama, J. Jiang,
and R. Saito, Appl. Phys. Lett. 88(2) (2009), p. 23109.
[392] Y.-M. Lin, C. Dimitrakopoulos, K.A. Jenkins, D.B. Farmer, H.-Y. Chiu, A. Grill, and Ph. Avouris,
Science 327(5966) (2010), p. 662.