Вы находитесь на странице: 1из 16

Chapter 2

Literature review
This chapter presents the review of theoretical and experimental studies of graphene
and functionalized graphene. After the discovery of graphene in 2004 (Novoselov et
al. 2004) the research on graphene and functionalized graphene has grown so large
that thousands of papers are published every year (Taghioskoui 2009) and it is
practically impossible to provide a comprehensive review. In this chapter we have
only discussed most relevant papers and review articles.

While attempting to understand the electronic properties of 3D graphite P.R. Wallace


explored the theory of graphene in 1947 (Wallace 1947). During this period band
structure was established which is called Slonczewski-Weiss-McClure (SWM) band
structure (McClure 1957, Slonczewski and Weiss 1958). With the help of this model
experimental data could be described successfully (Boyle and Nozieres, 1958,
McClure 1958, Spry and Scherer1960, Soule et al. 1964, Williamson et al. 1965, and
Dillon et al. 1977). As the SWM model is not able to explain many phenomena (e.g.
the interaction between graphene planes) this model has been revised many times in
the past (Rydberg et al. 2003). The first synthesis and characterization of monolayer
graphene was achieved by Andre Geim and Kostya Novoselov in 2004 (Novoselov et
al. 2004) and they received Nobel Prize in physics in 2010 for this. One of the
interesting aspects of the graphene problem is the fact that massless, chiral Dirac
formations are its low-energy excitations. Graphene is a gap less semiconductor. The
valance and conduction band energies are linear function of momentum. The speed of
electrons in graphene is constant just as the speed of photons is a constant c, with the
difference that in graphene Dirac fermions move with a speed 300 times slower than
the speed of light. The quantum mechanics of graphene’s electrons is identical to that
of relativistic particles with vanishingly small mass. Thus its free particle states are
chiral. In neutral graphene the chemical potential crosses the Dirac point exactly. This
particular dispersion takes place only at low energies. As a result, many of the unusual
properties displayed by QED show up in graphene as well but at much slower speeds
(Castro Neto et al. 2006, Katsnelson et al. 2006 and Novoselov et al. 2007)

Another interesting feature of the Dirac fermions is the fact that they are insensitive to
electrostatic potentials. This is due to the so-called Klein paradox which states that
Dirac fermions can be transmitted with probability 1 through a classically forbidden
region (Calogeracos and Dombey 1999, Itzykson and Zuber 2006). Dirac fermions
behave in an unusual way in the presence of confining potentials. This behaviour
manifests itself in the phenomenon of Zitterwegung, or a jerky movement in the
functioning of the waves. These electrostatic potentials are easily generated in
graphene by disorder. Since some disorder is inevitable in all materials, this has
generated great interest in trying to understand how this disorder affects the physics
and transport properties of graphene electrons. It was realized early in the research on
graphene that graphene would also show unusual mesoscopic effects (Peres et al.
2006, Katsnelson et al. 2007). These effects have their origin in the boundary
conditions required for the wave functions in mesocopic samples. These samples have
the various types of edges that graphene can have (Nakada et al.1996, Wakabayaashi
et al.1999, Peres et al. 2006, Akhmerov and Beenakker 2008). The zigzag and
armchair edges, which are the most studied edges, have extremely different electronic
properties. Zigzag edges sustain edge (surface) states and resonances that have never
been found in the armchair edges. When coupled to a conducting lead, the boundary
conditions of a graphene ribbon strongly affect its ability as a conductor. During this
condition the chiral Dirac nature of fermions in graphene can be used for applications
where one can control both the valley flavor of the electrons and also the so-called
valleytronics charge (Rycerz et al. 2007).

2.1 Properties of Graphene

2.1.1 Structural and Electronic Properties of Graphene


The atomic structure of isolated, single-layer graphene was studied by suspending
sheets of graphene between bars of a metallic grid and exposing them to transmission
electron microscopy (TEM) (Meyer 2007). The electron diffraction patterns that
resulted showed the typical and expected honeycomb lattice pattern of graphene. The
suspended graphene also showed a “rippling” of the suspended sheet of about one
nanometer amplitude. There could be two causes for this rippling. Either it was
intrinsic to graphene as a result of the instability of two dimensional crystals ( Geim
et al. 2007, Carlsson 2007, Fasolino et al. 2007) or it may be caused by extrinsic
element, the ever present dirt that can be seen in all TEM images of graphene. Atomic

resolution real-space images of isolated, single layer graphene were obtained on SiO 2
substrates (Masa et al. 2007, Elena et al 2007) by scanning the results of tunneling
microscopy. Graphene processed by using lithographic techniques is covered by photo
resistant residue. This residue may be the “adsorbates” observed in TEM images and
may be the cause of the rippling of suspended graphene. It was finally established that
the rippling of graphene on the SiO2 surface was caused by the conformation of

graphene to the underlying SiO2 and not to any intrinsic effect (Masa et al, 2007).
Graphene sheets in solid form show evidence of diffraction of 0.34mm
(2) in the graphite layering. This phenomenon also occurs in some single-walled
carbon nanostructures (Kasuya et al. 2002). A University of Manchester study,
published in the journal “Mesoscale and Nanoscale Physics” (Zan et al.2012) shows
that graphene can repair holes in graphene sheets on its own, when it is exposed to
molecules like hydrocarbons which contain carbon. When bombarded with pure
carbon atoms, the holes in graphene sheets are completely filled up, with the carbon
atoms snapping to the gaps and aligning themselves perfectly into hexagonal shapes.

Graphene differs from most conventional three-dimensional materials in the fact that intrinsic
graphene is a semi-metal or zero-gap semi-conductor. As stated earlier, the starting point for finding
the band structure of graphite is to understand the electronic structure of graphene. P.R. Wallace
realized as early as 1947 (Wallace 1947) that for the low energies near the six corners of the two-
dimensional hexagonal Brillouin zone, the E-K relation is linear. This leads to a zero effective mass for
electrons and holes (Charlier et al, 2008). This in turn leads to the fact that electrons and holes near
these six points, two of which are in equivalent, behave like relativistic particles as described by the
Dirac equation for spin ½ particles (Semenoff 1984, Avourius et al, 2007). Hence the electrons and
holes are called Dirac fermions (they are also sometimes referred as Graphinos) (Saul 2012) and the
six corners of the Brillouin zone are called the Dirac points (Charlier et al. 2008). The Brillouin zone of
graphene consists of two momentum positions (Dirac points) at the corners of the zone labeled as K
and K' respectively. We define their positions as follows,

2 2 2 2
= , , ′= ,− (2.1)
3 3 3 3
3 3

These Dirac points are presented on figure (2.1), where the Γ point is found at the
centre of the Brillouin zone and is non degenerate.

The tight binding method is based on the Hamiltonian of graphene being written as,
=− b +a (2.2)

ϯ ϯ
a b
n n+σi n n+σi

n,σi

Figure 2.1 Honeycomb lattice (left) of graphene and its Brillouin zone (right)
(figure taken from Castro et al. 2009)

Figure 2.2 The band structure of the graphene structure

Where t is the hopping parameter, an and bn+σi are the annihilation operators and aϯ and
bϯ are the creation operators corresponding to sublattices of graphene. After diagonalizing the
Hamiltonian equation, the operators are solved as follows:

1 , 1
= = (+)() (2.3)
+

Where N is the number of primitive unit cells. Substituting the annihilation operators
into (2.2) and applying the summations, we obtain the wavefunctions of graphene as
follows:
3

2
cos (2.4)
3 −
= − 1+2
2

Theoretically, the energy bands are derived by the eigenvalues

=± (2.5)

The equation (2.6) is responsible for the theoretical calculations of the band structure of graphene
materials.

=± 1+4
2 + 4cos⁡( )cos⁡( ) (2.6)
2 2 2

Figure (2.2) shows the tight binding calculated band structure from equation (2.6).
We observe that the band dispersion is conical at the Dirac points, unlike that in some
crystal materials which have a parabolic dispersion. The dashed line, which is called
the Fermi level, lies exactly at the intersection points. This phenomenon reveals to us
that graphene material is a gapless semiconductor.

2.1.2 Optical Properties of Graphene

The unique optical properties of graphene produce an unexpectedly high opacity for
an atomic monolayer. This opacity has a startlingly simple value. It absorbs πα ≈ 2.3%
of white light, where α is the fine-structure constant (Kuzmenko et al. 2008). This
simple value is a result of low-energy electronic structure of monolayer graphene.
This structure features electron and hole conical bands meeting each other at the Dirac
point. This is qualitatively different from more common quadratic massive bands
(Nair et al. 2008). Based on the Slonczewski–Weiss–McClure (SWM) band model of
graphite, the interatomic distance, hopping value and frequency cancel each other out
when the optical conductance is calculated in the Fresnel equations. This has been
confirmed experimentally, but the measurement is not yet precise enough to help us to
improve other techniques which are used for determining the fine-structure constant.
The band gap of graphene can be tuned from 0 to 0.25 eV (about 5 micrometer
wavelength) by applying voltage to a dual-gate bilayer graphene field-effect transistor
(FET) at room temperature (Zhang et al. 2009). The optical response of graphene
nanoribbons has also been shown to be tunable into the terahertz regime by an applied
magnetic field (Lui 2008). It has been shown that graphene/graphene oxide system
exhibits an electro chromic behavior, which allows
for the tuning of both linear and ultrafast optical properties (Kurum et al. 2011).
Recently, a graphene-based Bragg grating (one-dimensional photonic crystal) has
been fabricated and demonstrated its competence for excitation of surface
electromagnetic waves in the periodic structure using prism coupling technique
(Sreekanth et al. 2012). It has been further confirmed that this unique absorption could
become saturated whenever the input optical intensity rises above a threshold value.
This nonlinear optical behavior is called saturable absorption and the term given to
threshold value is saturation fluence. Graphene can be saturated quite easily under
strong excitation over the visible to near-infrared region, because of the universal
optical absorption and zero band gap. This property has relevance when we study the
mode locking of fiber lasers, where full band mode locking has been achieved by
graphene-based saturable absorber. It is because of this special property that graphene
has come to have wide application in ultrafast photonics. Moreover, the ultrafast
optical response of graphene/graphene oxide layers can be tuned electrically (Kurum
et al. 2011, Qiaoliang et al. 2009, Zhang et al. 2009). Furthermore, Saturable
absorption in graphene could occur at the Microwave and Terahertz band, owing to its
wideband optical absorption property.

2.1.3 Magnetic Properties of Graphene


Considerable interest has been aroused in the magnetism of nanographite particles for
some time. Enoki et al. (Enoki and Kobayashi 2005) have established that in
determining the magnetic properties of nanographite particles important role is played
by edge states of the adsorbed or intercalated species. Recently the magnetic
properties of graphene samples prepared by exfoliation of graphite oxide have been
studied. Studies have also been made of the conversion of nanodiamond, arc
evaporation of graphite and partial chemical reduction of graphene oxide (Matte 2009,
Wang 2009). All these studies have shown that there is a divergence between field-
cooled (FC) and zero-field-cooled (ZFC) data at 500 Oe. The studies also reveal that
the divergence almost disappears on the application of 1T. The magnetic properties of
graphene show us that dominant ferromagnetic interactions coexist along with
antiferromagnetic interactions. This was found to be true in all of the samples. It is not
possible to pin down exactly the origin of magnetism in the graphene samples
although it is established that defects and edge effects are likely to play a significant
role. Adsorption of benzene solutions of TTF and TCNE has a profound effect on the
magnetic properties of graphene. It is an ideal material for spintronics due to small
spin-orbit interaction and near absence of nuclear magnetic moments in carbon.
Electrical spin-current injection and detection in graphene was recently demonstrated
up to room temperature (Nikolaos et al. 2007, Sungiae et al. 2007, Megumi et al.
2007). Spin coherence length above 1 micrometre at room temperature was observed,
(Nikolaos et al. 2007) and control of the spin current polarity with an electrical gate
was observed at low temperature (Sungiae et al. 2007).

2.1.4 Transport Properties of Graphene

Experimental results from transport measurements show that graphene has a


remarkably high electron mobility at room temperature, with reported values in excess
2 −1 −1
of 15,000 cm ·V ·s (Geim and Novoselov 2007). The symmetry of the
experimentally measured conductance indicates that the mobilities for holes and
electrons should be nearly the same (Charlier 2008). The mobility is nearly
independent of temperature between 10 K and 100 K, (Novoselov et al. 2005,
Morozov et al. 2008, Chen et al. 2008) which implies that the dominant scattering
mechanism is defect scattering. Scattering by the acoustic phonons of graphene places
2 −1 −1
intrinsic limits on the room temperature mobility to 20,000 cm ·V ·s at a carrier
density of 1012 cm−2 (Chen et al. 2008, Akturk and Goldsman 2008). The
−6
corresponding resistivity of the graphene sheet would be10 Ω·cm. This is less than
the resistivity of silver, the lowest resistivity substance known at room temperature.
However, for graphene on SiO2 substrates, scattering of electrons by optical phonons
of the substrate is a larger effect at room temperature than scattering by graphene’s
2 −1 −1
own phonons. This limits the mobility to 40,000 cm ·V ·s (Chen et al. 2008).
Despite the zero carrier density near the Dirac points, graphene exhibits a minimum
2
conductivity of the order of 4e /h. The origin of this minimum conductivity is still
unclear. However, rippling of the graphene sheet or ionized impurities in the SiO 2
substrate may lead to local puddles of carriers that allow conduction (Charlier et al.
2008). Several theories suggest that the minimum conductivity should be 4e2/Πh,
2
however, most measurements are of order 4e /h or greater (Geim and Novoselov
2007) and depend on impurity concentration (Chen et al. 2008). Many recent
experiments have focused on the influence of chemical dopants on the carrier mobility
in graphene (Chen et al. 2008, Schedin et al. 2007). Schedin et al. experimented by
doping graphene with various gaseous species which were acceptors and donors. It
was found that the doped graphene could be restored to its undoped structure by
heating it gently in vacuum. This study found that even when chemical dopant
12 2
concentration is in excess of 10 cm there is no observable change in the carrier
mobility (Schedin et al. 2007). Chen, et al. doped graphene with potassium in ultra
high vacuum at low temperature. They found that potassium ions act as expected for
charged impurities in graphene, (Adam et al. 2007) and can reduce the mobility 20-
fold (Chen et al. 2008). The mobility reduction can be reversed by heating graphene
in order to remove potassium. Charge fractionalization is thought to occur in graphene
due to its two-dimensional property (Steinberg et al. 2008). It may therefore be a
suitable material for the construction of quantum computers using any ionic circuit
(Pachos 2009, Franz 2008). Charge carriers in graphene are massless Dirac fermions
that behave fundamentally differently than electrons in conventional semiconductors.
The most interesting factor is the very high quality of graphene lattice leading the
ballistic transport over micron length scales. Thus the electron transport is widely
studied in graphene (Saloriutta 2013).

2.2 Review of Experimental Work on Graphene


Starting from the 1970s single layers of graphite were grown out epitaxially on top of
other materials (Oshima and Nagashima 1997). This "epitaxial graphene" has a single
2
atomic thick hexagonal lattice structure of sp -bonded carbon atoms, which is found
in free-standing graphene. There is, however a significant charge transfer from the
substrate to the epitaxial graphene. In some cases, there is hybridization between the d
orbitals of the substrate atoms and π orbitals of graphene. This hybridization alters the
electronic structure of the epitaxial graphene significantly. Through the use of
transmission electron microscopy single layers of graphite were also observed within
bulk materials. This was particularly noticeable in soot which have been obtained
through chemical exfoliation. A number of experiments have also been conducted to
make thin films of graphite by mechanical exfoliation (Geim and Philip 2008). But
these experiments have not been able to produce graphene thinner than 50 to 100
layers during these years.

Graphene has been prepared using different techniques i.e, micromechanical cleavage
of graphite (Novoselov et al. 2004), the epitaxial growth on SiC substrates (Berger et
al. 2004), thermal exfoliation of GO at high temperatures (Schniepp et al. 2006,
Subrahmanyam et al. 2008), heating nanodiamond at different temperatures
(Andersson et al. 1998, Prasad et al. 2000) and Hummers method (Hummers et al.
1958). Hanns-Peter Boehm (Boehm, Setton and Stumpp 1994, Schniepp et al. 2006)
reported monolayer flakes of reduced graphene oxide. Graphite oxide reduction was
probably historically first method of graphene synthesis (Boehm et al. 1964). The first
attempt to produce individual graphene sheets by exfoliation dates to the work of
Brodie (Brodie 1859). Since then, and despite many attempts, (Staudenmaier 1898,
Hummers and Offeman 1958, Boehm and Scholtz 1965, Lueking et al. 2005,
Fukushima, Drzal and Annu 2003, Matsuo et al. 2003, Stankovich et al. 2006) large-
scale production of single graphene sheets has not been achieved.

Figure 2.3 Thin graphene “pancakes” deposited by “nano pencil” onto the silicon
wafer. The samples in the electron micrograph are magnified 6000× (left), like a silk
sheet, folded sheets of graphene on a silicon plate. The image was made with a
scanning electron microscope, magnified about 5000 times (middle), Thin graphitic
flakes on a surface of Si/SiO2 wafer (300 nm of SiO2, purple colour). The different
colours correspond to flakes of differing thicknesses, from ~100 nm (the pale yellow
ones) to a few nanometres (a few graphene layers – the most purple ones). The scale
is given by the distance between the lithography marks (200 μm) (right) (figures taken
from Geim & Philip 2008, Castro et al. 2006).
Single-layer graphene (SG) was first prepared by micromechanical cleavage from
highly ordered pyrolytic graphite (HOPG) (Novoselov et al. 2004). In this procedure,
a layer is peeled off the HOPG crystal by using scotch tape and then transferred on to
a silicon substrate. A chemical method to prepare single-layer graphene involves
reduction of single-layer graphene oxide (SGO) dispersion in dimethylformamide
with hydrazine hydrate (Park et al. 2009, Hummers and Offeman 1958). The Graphite
oxide (GO) is first prepared by oxidative treatment of graphite by employing
Hummers procedure, by the reaction of graphite powder (500 mg) with a mixture of
concentrated H2SO4 (12 ml) and NaNO3 (250 mg) in a 500 ml flask kept in an ice
bath. While stirring the mixture, 1.5 g of KMnO4 is added slowly and the temperature
brought up to 35°C. After stirring the mixture for 30 minutes, 22 ml of water is slowly
added and the temperature raised to 98°C. After 15 minutes, the reaction mixture is

diluted to 66 ml with warm water and treated with 3% of H2O2. Then suspension so
obtained is filtered to obtain a yellow-brown powder. This is washed with warm
water. GO readily forms a stable colloidal suspension in water and the suspension is
subjected to ultrasonic treatment (300 W, 35 KHZ) to produce single-layer graphene

oxide (SGO). The SGO (0.3 mg/ml) suspension in a H 2O+ N, N-dimethylformamide


(DMF) mixture (50 ml) is treated with hydrazine hydrate at 80°C for 12 hours (Sorella
and Tosatti 2001). This yields a black suspension of reduced graphene oxide (RGO)
in DMF/H2O. To make a stable dispersion of RGO, a further amount of DMF is
added to the suspension. It is useful to distinguish this single-layer material (RGO)
from the SG obtained by micromechanical cleavage of graphite or other means since
RGO may yet contain some residual oxygen functionalities. Gram quantities of
single-layer graphene have been obtained by a solvothermal procedure using sodium
and ethanol. Exfoliation of graphite in N-methylpyrrolidone or a surfactant/water
solution employing ultrasonication yields stable SG dispersions (Hernandez et al.
2008, Lotya et al. 2009). SG films are produced on the Si- terminated (0001) face of
single crystal 6H-SiC by thermal desorption of Si (Berger et al. 2004, Rollings et al.
2006, Emtsev et al. 2009). In this procedure, the substrates are subjected to electronic
bombardment in ultrahigh vacuum to 1000°C. This done to remove oxide
contaminants. The substrates are then heated to temperatures ranging from 1250 to
1450°C for 1-20 min. SG is prepared more conveniently using chemical vapor
deposition (CVD) by decomposing hydrocarbons on films or sheets of transition metal
such as Ni, Cu, Co and Ru (Reina et al. 2009). The graphene layers are grown on
different transition metal substrates by decomposing a variety of hydrocarbons such as
methane, ethylene, acetylene and benzene, the number of layers varying with the
hydrocarbon and reaction parameters. In these experiments, nickel (Ni) and cobalt
(Co) foils with thickness of 0.5 mm and 2 mm respectively were used as catalysts.

These foils were cut into 5 × 5 mm 2 pieces and polished mechanically and the CVD
process carried out by decomposing hydrocarbons around 800-1000°C. A nickel foil
was employed to carry out CVD passing methane (60-70 sccm) or ethylene (4-8
sccm). Along with this of high flow of hydrogen around 500 sccm at 1000°C are also
passed. This procedure was carried out for 5-10 minutes. With benzene acting as the
hydrocarbon source, benzene vapor diluted with argon and hydrogen was decomposed
at 1000°C for 5 minutes. Acetylene (4 sccm) and methane (65 sccm) were
decomposed at 800 and 1000°C respectively on a cobalt foil. The common feature of
all these experiments was that the metal foils were cooled gradually after the
decomposition.

An important method to prepare few-layer graphene (EG) is by thermal exfoliation of


GO at high temperatures (Schniepp et al. 2006, Subrahmanyam et al. 2008). In this
method, graphitic oxide (GO) is first prepared by the Staudenmaier method which is
as follows. A mixture of, sulfuric acid (10 ml) and nitric acid (5 ml) is taken in a 250
ml reaction flask cooled in an ice bath and graphite (0.5 g) is added under vigorous
stirring to the acid mixture. After the graphite powder is fully dispersed, potassium
chlorate (5.5 g) is added slowly over 15 min. The reaction mixture is stirred for 96 h
at room temperature. On completion of the reaction, the mixture is filtered and the
residue (GO) is washed in a 5% solution of HCl. GO is then washed repeatedly with
deionized water until the pH of the filtrate is neutral and then dried in vacuum at
60°C. GO so prepared (0.2 g) is placed in an alumina boat and inserted into a long
quartz tube sealed at one end. The sample is purged with Ar for 10 min, and then
quartz tube is quickly inserted into a tube furnace preheated to 1050°C and held in the
furnace for 10 minutes. The sample obtained for this procedure corresponds to the
few-layer graphene (EG). There is another method for preparing few-layer graphene.
In this method SGO make in water with hydrazine hydrate at the refluxing
temperature or by microwave treatment (EG-H) (Rao et al. 2009, Stankovich et al.
2007). Hydrazine hydrate (1 ml) is added to 100 ml of stable aqueous exfoliated
graphene oxide solution (1 mg/1 ml). Reflection is than carried out for 24 hours. The
reduced GO turns black and precipitates to the bottom of the flask. In the next step of
the experiment the precipitate is filtered and washed with water and methanol.
Ethylene glycol can be used instead of hydrazine hydrate use as the reducing agent in
the preparation of few-layer graphene (EG-H (G)). In this experiment a homogeneous
mixture of 25 ml of exfoliated graphene oxide and 2 ml of ethylene glycol in a 50 mL
PTFE-lined auto claver. This is then sealed and kept in an oven at a temperature of
170°C for 24 h under autogenously pressure. After this period of time it is allowed to
gradually cooled to room temperature. The product of this procedure is than washed
with water and ethanol. Another method of preparing graphene is by heating
nanodiamond in an inert or a reducing atmosphere. The effect of heating
nanodiamond at different temperatures has been studied by Enoki et al., (Andersson et
al. 1998, Prasad et al. 2000). Few-layer graphene (DG) can also be produced by the
annealing of nanodiamond at high temperatures in an inert atmosphere
(Subrahmanyam et al. 2008, Andersson et al. 1998).

In most studies, the starting material is graphite oxide (GO) produced through an acid
treatment of graphite. GO is then exposed to either a thermal or mechanical (i.e.,
ultrasonication) treatment to expand or to exfoliate it. Although nanoplates of a few
sheets have been generated by the solution approach (Stankovich et al. 2006, Du et al.
2005) the failure to produce single sheets by thermal expansion appears to be either
due to insufficient oxidation of graphite during the acid treatment or inadequate
pressure buildup during the thermal heat treatment stages. Schniepp et al. (Schniepp et
al. 2006) through an optimal combination of GO preparation and thermal treatment,
bulk quantities of functionalized single graphene sheets can be produced. The term
graphene first appeared to describe single sheet of graphite as one of the constituents
of graphite intercalation compounds (GICs); conceptually a GIC is a crystalline salt of
intercalant and graphene (Mouras et al. 1987). In graphite intercalation compounds,
the graphene layers either accept electrons from or donate electrons to the intercalated
species. It has now been generally accepted, however, that the description of the
electronic exchange in these compounds be specified from the standpoint of the
acceptor or donor properties of the intercalated species rather than from that of the
graphene layers, as customary with doped semiconductors. The term was also used in
early descriptions of carbon nanotubes as well as for epitaxial graphene and
polycyclic aromatic hydrocarbons.

Chiral Quantum Hall Effects: Considerable experimental efforts have been focused
on the electronic properties of graphene in order to understand the consequences of its
QED like spectrum. Some of the spectacular phenomena reported are (i) two new
chiral quantum hall effects (QHE’s) (ii) minimum quantum conductivity in the limits
of vanishing concentration of charge carriers (iii) strong suppression of quantum
interference effects (Geim & Novoselov 2007 and references there in)
Figure 2.4 Chiral quantum Hall effects. (a) The hallmark of massless Dirac fermions is QHE plateaux in σxy
at half integers of 4e2/h. (b) Anomalous QHE for massive Dirac fermions in bilayer graphene is more subtle:
σxy exhibits the standard QHE sequence with plateaux at all integer N of 4e2/h except for N = 0. The missing
plateau is indicated by the red arrow. The zero-N plateau can be recovered after chemical doping, which
shifts the neutrality point to high Vg so that an asymmetry gap (≈0.1eV in this case) is opened by the electric
field effect (c) Different types of Landau quantization in graphene. The sequence of Landau levels in the
density of states D is described by EN ∝ √N for massless Dirac fermions in single-layer graphene (d) EN ∝
√N (N–1) for massive Dirac fermions in bilayer graphene (e) The standard LL sequence EN ∝ N + ½ is
expected to recover if an electronic gap is opened in the bilayer (figure taken from Geim & Novoselov
2007).

2.3 Review of Functionalized Graphene


The absence of a band gap in graphene hinders its applications in microelectronic
devices, yet a suitable band gap can be tailored by chemical functionalization which
results into the tuning of electronic and magnetic properties for vast applications in
microelectronics. Several methods can be attempted for this, but surface chemical
functionalization is more attractive because graphene consists of only surface atoms.
In order to fulfill this purpose, attempts have been made to incorporate nitrogen,
halogens as well as oxygen for preparing several graphene derivatives (Bushan 2010,
Sluiter & Kawazowe 2003 and Sofo et al. 2007).

Hydrogenation of graphene: An interesting experimental work using scanning


tunneling microscopy, by Hornekaer et al (Hornekaer et al. 2006) demonstrates
clustering of hydrogen atoms on graphite surface. The work shows that there is
vanishingly small adsorption barrier for hydrogen in the vicinity of already adsorbed
hydrogen atoms. The question of the effect of defect on the electronic structure of
atomic hydrogen has been addressed by Duplock et al. (Duplock et al. 2007) within
the density functional theory. They show that the electronic gap state associated with
the adsorbed hydrogen is very sensitive to the presence of defect such as Stone-Wales
defect. The first theoretical description of functionalized graphene i.e. graphane was
reported in 2007 (Sofo et al. 2007). The electronic structure of hydrogen adsorbed on
graphene has been investigated using density functional theory (DFT) by Boukhvalov
et al. (Boukhvalov et al. 2008) and Casolo et al. (Casolo et al. 2009). Their results
support the use of graphane as a hydrogen storage material. Their work also shows
that the thermodynamically and kinetically favored structures are those that minimize
the sub lattice imbalance. Flores et al. have investigated the role of hydrogen-
frustration in graphane like structures using ab-initio methods and reactive classical
molecular dynamics (Flores et al. 2009). The stability trends in small clusters of
hydrogen on graphene have been discussed in a review (Roman et al. 2009). The
preparation of hydrogenated graphene was reported in 2009 by Geim’s group (Elias et
al. 2009). They have experimentally demonstrated for the first time that a single layer
of graphene can be hydrogenated to graphane which has been until now only a
theoretical material. The hydrogenation of graphene was effected by using a cold
hydrogen plasma for 2 hours. Based on electrical and structural characterization, the
formation of graphane was confirmed. In contrast to graphene, graphane exhibited an
insulating behaviour such that the resistivity grew by two orders of magnitude when
the temperature was decreased from 300 K to 4 K. The observed temperature
dependencies of resistivity of graphane could be fitted to a variable-range hopping

model. The carrier mobility changed from 14,000 cm2 V–1 s–1 down to values of

nearly 10 cm2 V–1 s–1 at liquid-helium temperatures. The quantum Hall plateaus
completely disappeared with only weak signatures of Shubnikov–de Haas oscillation
remaining in a magnetic field of 14 T. Graphene’s ambipolar field effect with the
neutrality point near zero-gate voltage has not been observed after its hydrogenation.
Graphane was found to be stable at room temperature for many days and could be
reconverted to graphene by annealing at 450°C in Ar atmosphere for 24 hours, albeit

the carrier mobility could be recovered to only about 3500 cm 2 V–1 s–1 (Srinivasan
and Saraswathi 2010). This process of fabrication of hydrogenated graphene is also
known as gas phase hydrogenation. There are few other methods of the fabrication of
hydrogenated graphene i.e. Liquid phase hydrogenation (Yang et al. 2012), Bottom up
approach (Wang et al. 2010) and pattering of hydrogenated graphene strips in a
graphene lattice (see Pumera and Wong 2013 and references cited there in).
Zhou et al. (Zhou et al. 2009) predicted a new ordered ferromagnetic state obtained by
removing hydrogen atoms from one side of the plane of graphane. Wu et al.
(Menghao et al. 2010) have reported implications of selective hydrogenation by
designing an array of triangular carbon domains separated by hydrogenated strips.
The electronic structure of the interface between graphene and the hydrogenated part
has been investigated by Schmidt and Loss (Schmidt and Loss 2010). They show the
existence of edge states for zigzag interface having strong spin orbit interaction. The
localization behaviour of disordered graphene by hydrogenation has been reported
within the tight binding formalism (Bang and Chang 2010). The potential of
hydrogenated graphene nanoribbons for spintronics applications has been investigated
within DFT, by Soriano et al. (Soriano et al. 2010). The defect and disorder induced
magnetism in graphene (including adsorbed hydrogen as a defect) has been studied by
Yazyev and Helm mostly using Hubbard Hamiltonian and DFT (Yazyev 2010,
Yazyev 2008, Yazyev and Helm 2007).

Alzahrani and Srivastava presented ab initio calculations to investigate the geometric


and electronic properties of graphene. They determined that the chair like
conformation is the most energetically stable structure for graphane (Alzahrani and
Srivastava 2010). Rozhkov et al. discussed electronic properties of mesoscopic
graphene based structures and found that it is an attractive candidate for
microelectronic and micromechanical applications (Rozhkov et. al 2011). Quantum
dots (QD's) or artificial atoms are one of the most intensely studied systems
(Chakraborty et. al 1962, 1999). Quantum dots, discovered in the 1980s (Ekimov et.
al 1981), are semiconductors that contain a size and shape-dependent band gap. These
have been promising structures for applications that range from computers, LEDs,
solar cells and lasers to medical imaging devices. Further confinement toward
graphene QDs turns out to be a challenge that triggered intense studies of large
graphene dots (Singh et. al 2010). The larger QDs of graphene hold promise for
spintronics (Trauzettel et. al 2007), although their small gaps limit optoelectronics
applications. Theoretical investigations have also been carried out for a single
hydrogen defect on graphane sheet using GW method (Leb`egue et al. 2009), and for
one and two vacancies in graphane (Pujari and Kanhere 2009) using DFT. It has been
shown that interaction between adatoms in hydrogenated graphene is long range and
its nature is dependent on which sublattice the adatoms reside (Shytov, Abanin and
Levitov 2009). The result suggests that the adatoms tend to aggregate. A very recent
work explores a formation of quantum dots as small island of graphene in graphane
host (Singh, Penev and Yakobson 2010). Hydrogen adatoms such as resonant
scatterers can strongly enhance the low energy density of states in graphene (Shengjun
et al. 2012). Shengjun et al. studied the impact of impurities on electronic screening.
They found that, this bad metal behaviour manifests the dynamic polarization function
and can be directly measured by means of electron energy loss spectroscopy. It has
been noticed that all information on the structure and properties of graphane originates
from theoretical studies as there has not yet been any successful fabrication of pure
graphane (Pumera and Wong 2013).
Applications of graphane: Since graphane is a wide band gap semiconductor in
contrast to pristine graphene (which is a semi metal), it can be used in electronic
devices such as transistors. The stability and catalytic activity of graphane may be
improved by anchoring metal catalyst on its support. Due to the presence of C-H bond
graphane becomes an activated precursor for further chemical medications and thus
permits controllable functionalization path ways which are not available to graphene.
Hydrogenated graphene is found to be ferromagnetic in nature and it displays tunable
band gap depending on the extent of hydrogenation. It is also a potential candidate for
hydrogen storage due to its property of reversible hydrogenation.
Fluorination of graphene: The fluorination of graphene has attracted considerable
2
attention because it increases the reactivity of sp bonds and opens a wide range of
modifications. It enables tuning of band gaps and opens up possibilities as a dopant
for usage in electronics. Many experiments have shown dramatic changes in the
electronic and structural properties of graphene by increasing the percentage of
fluorine (Wang et al. 2012, Junkermeier et al. 2013).

Вам также может понравиться