Você está na página 1de 93

Density Functional Theory Calculations for

the MonolayerBilayer Graphene


Heterojunctions

Masterarbeit aus der Physik

Vorgelegt von
Daniel Branski
25.06.2014

Lehrstuhl fr theoretische Festkrperphysik


Friedrich-Alexander-Universitt Erlangen-Nrnberg

Betreuer: Prof. Dr. Oleg Pankratov


Contents

1 Introduction 5

2 Theory 9
2.1 The basics of Density Functional Theory . . . . . . . . . . . . . . . . 9
2.1.1 Hohenberg-Kohn-Theorems . . . . . . . . . . . . . . . . . . . 10
2.1.2 The Kohn-Sham Scheme . . . . . . . . . . . . . . . . . . . . . 11
2.1.3 Local Density Approximation . . . . . . . . . . . . . . . . . . 12
2.1.4 Projector Augmented-Wave Method . . . . . . . . . . . . . . . 13
2.2 Introduction to Graphene . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.1 Graphene Monolayer . . . . . . . . . . . . . . . . . . . . . . . 15
2.2.2 Graphene Bilayer . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Graphene Nanoribbons . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.1 Zigzag Termination . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.2 Armchair Termination . . . . . . . . . . . . . . . . . . . . . . 26
2.4 Bilayer Graphene Nanoribbons . . . . . . . . . . . . . . . . . . . . . . 27

3 Density Functional Theory Calculations for Graphene systems 29


3.1 Extended Graphene . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.1.1 Graphene Monolayer . . . . . . . . . . . . . . . . . . . . . . . 30
3.1.2 Graphene Bilayer . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 Constrained Graphene . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.1 Zigzag Termination . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.2 Armchair Termination . . . . . . . . . . . . . . . . . . . . . . 49
3.3 MonolayerBilayer Graphene junctions . . . . . . . . . . . . . . . . . 57
3.3.1 Junction with Zigzag Termination . . . . . . . . . . . . . . . . 57
3.3.2 Junction with Armchair Termination . . . . . . . . . . . . . . 77

4 Conclusion 85

References 87

3
1. Introduction

Graphene is a two dimensional honeycomb lattice of carbon atoms which is a basis


building block of the various carbon allotropes. These are graphite (3D), carbon
nanotubes (1D) and fullerenes (0D). Graphene was initially considered in theoretical
investigations [1] as a convenient starting point for "real" carbon-based structures,
but itself it was believed to be unstable and, therefore, non-existent. Yet in 2004
Geim and Novoselov [2] successfully isolated single graphene layers. This discovery
triggered an explosion of the research activity. Theoretically and experimentally, the
features of mono-, bi- and many-layer graphene and many systems around graphene
were studied. Two of them, the free standing nanoribbons and the nanoribbons
placed on the graphene monolayer are investigated this work. These systems realize
the monolayerbilayer junction, a novel solid state heterostructure.
In practice, such structures can be manufactured using epitaxial graphene grown
on SiC. It is well known that the first graphene layer on a silicon terminated SiC
surface is covalently bound the the substrate and does not possess the electronic
structure of the free standing graphene. This layer is usually called the buffer layer
which supports the second graphene layer (epitaxial monolayer graphene) showing
the genuine graphene properties [3]. Yet, there is a possibility to "activate" the buffer
layer such that it regains its graphene electronic structure by the intercalation of hy-
drogen between the SiC surface and the buffer layer. Thereby, the covalent bonds are
broken. The dangling bonds of the silicon atoms are saturated by the intercalated
hydrogen and the former buffer layer is pushed away from the surface restoring its
graphene properties [4]. This system is called quasi free standing graphene. Analo-
gously, quasi free standing bilayer graphene can be produced.
There is a crucial difference between epitaxial monolayer graphene, quasi free
standing graphene and quasi free standing bilayer graphene and its free standing
equivalents which are discussed in this thesis: the location of the Fermi level. While
the Fermi level lies at the Dirac points in the free standing systems, it is located above
the Dirac points in epitaxial graphene and below in quasi free standing graphene and
quasi free standing bilayer graphene [4, 5]. Following the nomenclature of semicon-
ductors, they are called n-type and p-type systems, respectively. If one considers
the junction between epitaxial graphene and quasi free standing bilayer graphene,
the Dirac points of the two structures lie at different energies. This property is not
covered by the simple model which is used in this work. Here, the Fermi level is

5
1. Introduction

always set to zero and coincides with the location of the Dirac points.
Graphene is one, if not the most, promising candidate to replace silicon in elec-
tronic devices and boost their performance significantly [6]. As pure graphene is
a zero band-gap semiconductor and a band gap is necessary for electronic devices,
nanoribbons and bilayer are of special interest because of their possibility to have
a band gap. Monolayerbilayermonolayer junctions are assumed to be used as
switches by applying a gate voltage in the bilayer region.
The two simplest edge geometries at a monolayer-bilayer interface or a terminated
graphene lattice are called zigzag and armchair termination based on their geometri-
cal shape. The properties of these edges are very different as will be discussed later.
Irregular and restructured edges are also possible and common in nature but require
the use of much larger unit cells for which density functional theory calculations are
impractical.
This thesis is organised as follows: Firstly, a brief description of density functional
theory is given, including sections on the work of Hohenberg, Kohn and Sham, the
local density approximation and the projector augmented-wave method. Secondly,
monolayer and bilayer graphene as well as zigzag and armchair terminated nanorib-
bons are described analytically. Sec. 3 starts with reproducing results for monolayer
and bilayer graphene with ab initio methods. Then, zigzag and armchair nanoribbons
are investigated. Finally, systems consisting of a graphene layer and a nanoribbon
on top are discussed as a model for the monolayer-bilayer and monolayer-bilayer-
monolayer interface.
The main goal of this thesis is to understand the electronic structure of graphene
monolayerbilayer junctions. Unfortunately, although the junction between epitaxial
monolayer graphene and quasi free standing graphene is of special interest, this
system is too large to be treated properly by density functional theory calculations.
Hence, only the free standing equivalent is treated: a junction between monolayer
and bilayer graphene. While it fails to reproduce the different Fermi levels relative to
the Dirac points, it may provide some insights. Moreover, the free standing system
itself is of interest and the focus of much research activity [7, 8].
Zigzag and armchair terminated junctions differ substantially in their electronic
structure. Localized states exist at zigzag but not at armchair edges. These localized
states correspond to flat bands near the Fermi level and are well understood in the
context of zigzag nanoribbons. In a monolayerbilayer junction, they induce localized
states on the continuous graphene layer. The properties of these induced states
depend on the geometry of the junction. Due to the AB stacking, two configurations
are possible. In both cases, bulk states near the Dirac points have a finite amplitude
in the monolayer region and on both layers in the bilayer region.
This is in contrast to junctions with armchair termination. While no localized
or induced states exist, bulk states near the Dirac point can be characterized in
monolayer and bilayer states. The bilayer states show a finite amplitude on the
continuous and the terminated graphene layer. Their curvature is correlated to the

6
band gap of the corresponding free standing armchair nanoribbon of the same width
as the bilayer region. The monolayer states are located on the continuous layer in
the monolayer and the bilayer region without being significantly disturbed by the
presence of the second graphene layer which, in our case, is an armchair nanoribbon.
The two monolayer bands touch at the Fermi energy.
Additionally, both systems have been investigated without passivating the ter-
minated graphene layer with hydrogen. The armchair terminated junction does not
show considerable differences except for the additional bands near the Fermi level
which correspond to the dangling bonds and the fact that the monolayer bands touch
above the Fermi energy. In one case, the band structure does not show monolayer
and bilayer states but four bands which states extend over both graphene layers
and features a band gap. The zigzag terminated junction, however, behaves differ-
ently. The presence of the dangling bond at the edge leads to an energetic raise
of the localized -states. The amplitude of this raise depends heavily on the edge
geometry.

7
1. Introduction

8
2. Theory

2.1 The basics of Density Functional Theory


In the following, the basics of density functional theory which includes the Hohenberg-
Kohn theorems [9] and the Kohn-Sham equations [10], the local density approxima-
tion and the projector augmented-wave method [11] are summarized. They build
the basis of the used computational implementation of density functional theory, the
Vienna Ab Initio Simulation Package (VASP) [1214].

To describe the electronic structure of a solid, one usually has to solve a quantum
mechanical interacting many-body system which is, in most cases, unsolvable even
with modern computers. Using the Born-Oppenheimer approximation and treating
only valence electrons and ions (instead of all electrons and nuclei) simplifies of
the problem enormously, but it still remains unsolvable. The Born-Oppenheimer
approximation allows to separate electron and ion degrees of freedom because the ions
are typically 5 orders of magnitude heavier than the electrons. This approximation
breaks down in rare special cases only and it shall be used throughout this thesis.
An early method to describe electron systems was the Thomas-Fermi model which
is the precursor of the modern DFT. A system of electrons bound by one or many
nuclei is described by the non-relativistic Schrdinger equation:
 
H(r) = T + Vne + Vee (r) = E(r). (2.1.1)

N
Pe N
Pe N
Pe
1
Here T = Ti = p2
2m i
is the kinetic energy operator, Vne = v(r i ) the ex-
i=1 i=1 i=1
ternal potential, which describes the interaction with a crystal lattice and Vee =
P
i<j w(r i , r j ) the electron-electron interaction potential. Within the Thomas-Fermi
model, the kinetic energy density T [n(r)] is approximated by the kinetic energy of a
uniform, non-interacting electron gas taken at a given density n(r). With the Hartree
expression for the electron-electron energy plus the contribution of the external po-
tential one arrives at the Thomas-Fermi energy functional ET F [n(r)]. Minimization
of ET F [n(r)] produces the ground state density distribution and the ground state
energy. Unfortunately, this method fails to describe the shell structure of atoms and
it cannot explain the existence of molecules.

9
2. Theory

These deficits are removed in the modern DFT which generalizes the Thomas-
Fermi approach in such a way that it becomes a (formally) exact theory of an in-
homogeneous many-body problem. The DFT calculations are based on the Kohn-
Sham-equations, a set of equations which have to be solved iteratively. They have
their roots in the groundbreaking work of Hohenberg, Kohn and Sham published
in 1964 [9] (Hohenberg and Kohn) and 1965 [10] (Kohn and Sham). The first pa-
per states that the ground state energy is a functional of the ground state density
and both can be found by minimizing the energy with respect to the density if the
correct functional was known. The second work proposes a way to construct the
density from non-interacting orbitals (Kohn-Sham wave functions) and leads to the
Kohn-Sham equations. Because of their importance, both parts are discussed in the
following, based on the description given in [15].

2.1.1 Hohenberg-Kohn-Theorems
The first theorem is as follows:
Theorem 2.1.1 The electron ground state density n(r) determines the external po-
tential to within an additive constant and, therefore, the ground state energy E0 =
E0 [n].

Proof Let H = T + W + V be the Hamiltonian of the system. T and W are the


same for all possible potentials V V. All ground state wave functions |i i build
a subspace V = {|i i|i 1, ..., q} and each wave function n |i i yields a density
o
ni (r) = h|n(r)|i which are combined in the set Nv = ni | i = {1, ..., q} .
Now, we consider two different potentials V , V 0 that differ by more than a constant.
Otherwise they are considered equivalent as they do not lead to different physical
systems. May further be |i V and | 0 i V 0 obeying
 
H|i = T + W + V |i = E|i (2.1.2)
 
H 0 | 0 i = T + W + V 0 | 0 i = E 0 | 0 i (2.1.3)
If we assume |i = | 0 i, we obtain by subtracting the equations (2.1.3)
(V V 0 )|i = (E E 0 )|i, (2.1.4)
which is in contradiction to the assumption that the potentials differ by more than
a constant. Therefore, all ground state wave functions of different potentials are
different. Further, may n(r) and n0 (r) be the densities associated with |i and | 0 i,
respectively. By Ritz principle [15], the following equations hold:
Z
0 0
E = h|H|i < h |H| i = E + 0
d3 r(V V 0 )n0 (r) (2.1.5)
Z
E 0 = h 0 |H 0 | 0 i < h|H 0 |i = E + d3 r(V 0 V )n(r) (2.1.6)

10
2.1. The basics of Density Functional Theory

If one assumes n(r) = n0 (r), this results in the inequality E + E 0 < E + E 0 . Hence,
the two densities must be different and, as different potentials lead to different den-
sities, there is a one-to-one map between V and N = {NV |V V}. 

The energy is given by


Z
E[n] = h[n]|T + W + V |[n]i =: FHK [n] + d3 r n(r) V (r). (2.1.7)

The Hohenberg-Kohn functional was defined in the last step. It only depends on
the interaction potential W , which is the Coulomb interaction between electrons, if
one is interested in the electronic structure of matter. The interaction with nuclei,
molecules or crystal lattices is determined by V . Because FHK [n] is the same for all
such systems, it is called universal functional. In case of a non-degenerate ground
state, the density uniquely determines the wave function and, hence, all expectation
values of the system.

The second theorem is


Theorem 2.1.2 The ground state energy E0 can be determined by a variational
principle: E0 = min E[n]
nN

The conditions on N have been the focus of research activity for many years, not
only because of its own purpose but also for the validity of computational imple-
mentations. Both theorems have also been extended to more general systems which
include more effects, i.e, spin polarization and ensemble states.
The Hohenberg-Kohn theorems are mathematical proofs that the ground state
energy is a functional of the electron density but, as the functional FHK remains
unknown, it is impractical for actual calculations.

2.1.2 The Kohn-Sham Scheme


Kohn and Sham proposed the following approach:
Theorem 2.1.3 For each interacting system A with energy functional as given in
Eq. (2.1.7), there exists a non-interacting system B, described by the Hamiltonian
H = TB + VB with an appropriate VB , which yields the same ground state density
nA (r) = nB (r).
The first Hohenberg-Kohn theorem guarantees that the uniqueness of VB for a given
density nB (r). The problem which remains is to find a suitable potential VB . As
the auxiliary system is non-interacting, the density can be constructed from single-
particle orbitals |i i by
N

X
n(r) = hi |n(r)|i i. (2.1.8)
i=1

11
2. Theory

Starting from Eq. (2.1.7), one can rewrite the energy functional

E[n] = T [n] + W [n] + V [n] = TB [n] + VH [n] + V [n] + Exc [n] (2.1.9)

with Exc [n] = T [n] TB [n] + W [n] VH [n] the exchange correlation functional.
One estimates the interacting kinetic energy by the non-interacting kinetic energy
TB = 21 2 and the interaction by the classical interaction VH [n] = 21 n(r 1 )n(r 2 )
R
|r 1 r 2 |
.
The difference made by this substitution is put in Exc in which all unknown quantities
are united. The auxiliary potential VB can be estimated by

VB [n] = V [n] + VH [n] + Exc [n]. (2.1.10)

Because of the second theorem of Hohenberg and Kohn, the variation E[n] vanishes
and leads to Euler-Lagrange equations which can be rewritten in the following form:

n(r 0 )
!
1 2 Z
+ v(r) + + vxc (r) i (r) = i i (r), (2.1.11)
2 |r r 0 |

with the condition


Z N Z
2
X
dr |(r)| = dr n(r) = N (2.1.12)
i=1

and the definition


Exc [n]
vxc = (2.1.13)
n
are the Kohn-Sham equations. The i and i are the Kohn-Sham orbitals and eigen-
values, respectively. Because the equations (2.1.11) and (2.1.13) depend on the den-
sity n(r) in (2.1.12) which itself depends on the Kohn-Sham orbitals, the equations
have to solved iteratively. If the exact expression for Exc was known, the resulting
ground state energy and density would be exact. As this is not the case, one has to
find a good approximation. The simplest and widely used approximation is briefly
presented in the following section.

2.1.3 Local Density Approximation


The local density approximation (LDA) is the simplest way to estimate the unknown
exchange-correlation energy contribution in the Kohn-Sham equations. The basic
assumption is that Exc can be approximated locally by the sum of the exchange
energy Ex and the correlation energy Ec of the homogenous electron gas of the same
density. The homogenous electron gas is a system of N electrons in a volume V with
a homogenous positive background charge which keeps the whole system neutral.
One takes the limit N and V with N V
= n = const. Ex is known exactly
in this model:
3 3 1/3 Z
 
Ex = dr n4/3 (r). (2.1.14)
4

12
2.1. The basics of Density Functional Theory

Unfortunately, the correlation energy can be found analytically only in the limits of
very high or very low densities. Yet, it was numerically calculated with quantum
Monte-Carlo techniques and parametrized as a function of n.
In this work, quantum Monte-Carlo results from Ceperly and Alder [16, 17] were
used, parametrized by Perdew and Zunger [18].
Theoretically, the LDA is only valid for systems in which the electron density
varies slowly. In practice, however, LDA calculations provide accurate results even
for very inhomogeneous systems like single atoms. The reason for this is partly given
by the fact that the correlation energy is usually overestimated while the exchange
energy is usually underestimated and these two errors tend to cancel [19].
Though many features are described accurately by LDA, systematic errors which
occur are too small lattice constants and band gaps. Therefore, whenever band gaps
and atomic distances are mentioned in this work, the true experimental value is likely
to be larger by a few percent.

2.1.4 Projector Augmented-Wave Method


The projector augmented-wave method was developed by P. E. Blchl in 1994 [11] as
a generalization of the linear augmented-plane-wave method and the pseudopotential
method. Both are used to avoid the treatment of the wave functions near nuclei,
which oscillate quickly due to the very deep Coulomb potential. In a plane wave
basis set, this requires a very high energy cutoff, which increases computational
effort without being necessary for the description of the valence states of interest. In
this section, a wave function has to be understood as the Kohn-Sham one-electron
function and not the many-electron wave function. In the following, the projector
augmented-wave method is summarized as it is described in [11, 20].
Blchls idea is to map the Hilbert space of all physical (all-electron, AE) wave
functions which are orthogonal to the core states on a so-called pseudo (PS) Hilbert
space.
|i = T |0 i, (2.1.15)
with T = 1+ R TR and R the atom sites. The operator TR acts only within a sphere
P

of radius R around each atom, the augmentation region. This means that |i and
|0 i coincide outside the augmentation region. Inside this region one can choose a
computationally convenient set of PS wave functions |0i i, which are orthogonal to
the core states and otherwise complete. These are used as the basis of PS Hilbert
space and they are mapped on a basis of (AE) wave functions |i i by
|i i = (1 + TR )|0i i, (2.1.16)
where i is a short hand for the atom site R, the shell number n and the angular
momentum (l, m). He proposes to use the solutions of the radial Schrdinger equation
as |i i. The linear operator T can be expressed as
(|i i |0i i)hp0i |.
X
T =1+ (2.1.17)
i

13
2. Theory

hp0i | are called projector functions which fulfill the condition hp0i |0j i = ij and can be
expressed most generally as
1
hp0i | = ({hfk |0l i}ij )
X
hfj |. (2.1.18)
j

Here, the hfi | are an arbitrary and linearly independent set of functions.
To calculate the expectation value of an operator A, h|A|i has to be evaluated.
To avoid the continual transformation between |i and |0 i, the operator A can be
transformed to its respective pseudo-operator A0 , which expectation value is eval-
uated by h0 |A0 |0 i. Depending on if the operator is quasi-local or nonlocal, the
transformation is done by
A0 = T AT (2.1.19)
or

A0 = T AT + A (2.1.20)

with

|p0i i (hi | h0i |) A 1 |0j ihp0j


X X
A =
i j
 
+ 1 |p0j ih0j | A (|i i |0i i) hp0i |,

respectively (for details see Ref. [11]).

14
2.2. Introduction to Graphene

Figure 2.1: The honeycomb structure of the graphene monolayer.


a1 and a2 are the two lattice vectors and 1 and 2 denote the two
carbon atoms per unit cell. The primitive unit cell is framed by the
lattice vectors and the dotted line.

2.2 Introduction to Graphene


In the following, the graphene monolayer and bilayer (in AB stacking) are briefly
discussed as they built the fundamental structures of nanoribbons which are inves-
tigated in more detail in the sections 2.3 and 3.3.

2.2.1 The extended Graphene Monolayer


The carbon atoms in graphene are sp2 -hybridized and form a honeycomb lattice
(Fig. 2.1). The interatomic distance is a 1.42 A. The three sp2 -orbitals build -
bonds to the three nearest neighbors while the remaining perpendicular pz -orbitals
constitute a delocalized -bonding part of the electron spectrum. The crystal struc-
ture is described by the primitive translation vectors
!
3 1
a1 = a , (2.2.1)
2 2
!
3 1
a2 = a , (2.2.2)
2 2

where a = 2.46 A is the lattice constant. The atomic positions are:

r 1 = (0, 0) (2.2.3)
!
1 1 1
r 2 = a1 + a2 = a , 0 (2.2.4)
3 3 3

The associated reciprocal lattice, shown in Fig. 2.2, is hexagonal lattice rotated by

15
2. Theory

Figure 2.2: Reciprocal space of Graphene: The Brillouin zone is a


hexagon. The points of symmetry are , K and M .

90 compared to the real space lattice.


!
4 1 3
b1 = , (2.2.5)
a 3 2 2
!
4 1 3
b2 = , (2.2.6)
a 3 2 2
The Brillouin zone has a honeycomb shape with the -point at its center. Only two
of the six edge points are inequivalent and they are usually called K and K 0 [21].
The very special property of graphene, which is a linear relation between the crystal
momentum k and energy for low energy excitations, takes place at these two specific
points in momentum space.
As can be seen from a simple tight binding model [21] and from DFT calculations
(see Sec. 3.1.1, p. 30), the low energy excitation states can be described by the Dirac-
Weyl equation.
This equation describes the massless ultra-relativistic particles with spin 12 (e.g.
neutrinos) with a linear relation between energy and momentum. It has the following
form:
0 ix + y 0 0

!
i 0 0 0
x y
H = ~vF . (2.2.7)

0 0 0 ix + y 0


0 0 ix y 0
Here, ~ is the Planck constant and the constant vF is called the Fermi velocity.
and 0 are bispinors and describe the electronic states at K and K 0 , respectively.
Because of this connection to ultra-relativistic physics, graphene is of great in-
terest not only for itself and its possible use in electronic devices but also as a model
system for particle physics and problems in quantum gravity. The linear dispersion
relation causes a linear behavior of the density of states at the Fermi energy Ef at
which it is exactly zero.

16
2.2. Introduction to Graphene

Figure 2.3: Two graphene layers in AB stacking: The primitive


unit cell, framed by the lattice vectors and the dotted line, contains
the four carbon atoms A1, B1, A2 and B2. A and B denote the
sublattice and 1 and 2 refer to the top and bottom graphene layer,
respectively.

2.2.2 The extended Graphene Bilayer


The bilayer graphene consists of two coupled graphene monolayers. The unit cell is
given by the following lattice vectors:
!
3 1
a1 = a , (2.2.8)
2 2
!
3 1
a2 = a , . (2.2.9)
2 2

a = 2.46 A is, again, the lattice constant. Instead of two atoms, the unit cell of
bilayer graphene contains four at the positions:
 
r A1 = a 0, 0, 0 (2.2.10)
!
1
r B1 = a , 0, 0 (2.2.11)
3
!
1
r A2 = a , 0, 1.42 (2.2.12)
3
2
 
r B2 = a , 0, 1.42 . (2.2.13)
3

The third component takes care of the interlayer distance d = 3.5 A. Labelling the
atoms of sublattice A and B A2 and B2 in the lower layer and A1 and B1 in the
upper, respectively, is most common in the literature. One can see that the atoms A2
and B1 are on top of each other. Hence, they are often referred to as "dimer" sites,

17
2. Theory

while the remaining atoms are called "non-dimer" sites. The associated reciprocal
lattice is, as the lattice vectors are the same, the same as for the monolayer (cf.
Fig. 2.2, p. 16):
!
4 1 3
b1 = , (2.2.14)
a 3 2 2
!
4 1 3
b2 = , (2.2.15)
a 3 2 2

The first Brillouin zone is a hexagon, where two of the six edges are inequivalent
which are called K and K 0 [21]. In contrast to monolayer graphene, where the
dispersion relation was linear at K and K 0 , the dispersion is quadratic. Due to the
larger unit cell, there are four bands of which two touch at the Fermi level.
A simple tight-binding model with intralayer hopping parameter 0 , interlayer
hopping at the dimer sites with parameter 1 and interlayer hopping between the
off-dimer sites with 3 leads to a Hamiltonian HK for the vicinity of K [21]:

0 ~vF k 0 3~3 ak
0 0

~vF k 1

HK = , (2.2.16)

0 1 0 ~vF k
3~3 ak 0 ~vF k 0

with k = kx + iky and k the complex conjugate of k. One can write an effective low
energy Hamiltonian if 3 , ~vF k  1 [21]:
(~vF k)2

0 1
HK = (~vF k)2 . (2.2.17)
1
0

Hence, electrons behave like massive particles near the Fermi energy due to the
parabolic band shape which is in contrast to the linear dispersion of monolayer
graphene. The parabolic behavior leads to a constant density of states for energies
very close to the Fermi energy.
Another important property of bilayer graphene is the possibility to from a band
gap by applying a gate voltage perpendicular to the bilayer sheet. Depending on the
voltage bands up to 300 meV have been observed.

18
2.3. Graphene Nanoribbons

2.3 Graphene Nanoribbons


When a graphene sheet is cut such that it forms a narrow strip, the resulting structure
is called graphene nanoribbon. Alternatively, one can think of a nanoribbon as
an unzipped carbon nanotube which is indeed a possible way of their fabrication
[22, 23]. Depending on the cutting direction, the nanoribbon has either zigzag edges
or armchair edges or a mixture of both, which is commonly referred to as a general
edge. Figures 2.4, 2.5 and 2.6 show the three types of nanoribbons and also define
the unit cells in case of zigzag and armchair nanoribbons.
Due to the constrained geometry, two new effects arise in nanoribbons. Firstly,
the sp2 -orbitals at the edge become dangling bonds which may or may not be satu-
rated. In investigations within the tight binding framework [24] or the Dirac-Weyl
equation [25], it is usually assumed that the dangling bonds are passivated e.g. by
hydrogen atoms and do not contribute to the electronic structure near the Fermi
level. While it is impossible to include sp2 dangling bonds in a Dirac theory, they
were included in a tight binding calculation [26]. Later in this work, the difference
between hydrogenated and non-passivated edges will be investigated with ab initio
method for zigzag and armchair terminations.
Secondly, the atomic positions will change due to the geometric termination of
the system. This effect is often neglected in tight binding investigations. Within
the Dirac-Weyl formalism, it cannot be implemented. It is known, that the carbon-
carbon distance has a crucial effect on the -orbit coupling [27]. The differences
between the ideal geometry and a relaxed system will also be discussed.

2.3.1 Zigzag Termination


It is well known from theoretical as well as experimental studies that zigzag termi-
nated graphene layers [25,28] and graphite steps [29,30] feature localized edge states.
For simplicity, the lattice constant is set to 1 throughout this section.
These states decay exponentially into the bulk with the penetration length being
a function of the k vector along the edge. Fig. 2.4 shows the unit cell of a zigzag
nanoribbon of width M, where M denotes the number of zigzag lines along the
periodic direction of the ribbon. As nanoribbons are one dimensional systems, the
Brillouin zone (BZ) is one dimensional and can be understood by projecting the bulk
graphene BZ on the proper axis . This is illustrated in Fig. 2.7. K and K 0 are located
at k = 23
and k = 2
3
, respectively.

19
2. Theory

Figure 2.4: Schematic structure of a zigzag nanoribbon. The edges


are denoted by bold lines and the unit cell is framed by the rectangle.

Figure 2.5: Schematic structure of an armchair nanoribbon. The


edges are denoted by bold lines and the unit cell is framed by the
rectangle.

20
2.3. Graphene Nanoribbons

Figure 2.6: Schematic structure of an irregular or general nanorib-


bon. The edges are denoted by bold lines. The structure is periodic
in the x-direction.

Figure 2.7: Brillouin zone of zigzag nanoribbons.

21
2. Theory

Electronic structure investigation using the Dirac-Weyl equation


The following part is reproduced from [25]. The Dirac-Weyl Hamiltonian is given
by:
0 ix + y 0 0

i 0 0 0
x y
H = ~vF ,

(2.3.1)
0 0 0 ix + y


0 0 ix y 0
where ~ is the Planck constant and vF the Fermi velocity in graphene. In this
formalism the electrons are described by a bispinor

A (x, y)

(x, y)
(x, y) = B0 , (2.3.2)

A (x, y)
B0 (x, y)

where the two upper components belong to K and the lower to K 0 . Because the
system is periodic in the x-direction, each component can be written as
(0) (0)
i = eikx x i (y). (2.3.3)

kx is measured from K (K 0 ), not from the center of the BZ. From now on, only
the upper part, which belongs to K, will be discussed, as the results for K 0 can be
derived by substituting x x or kx kx , respectively.
The Dirac-Weyl equation for K is
! !
0 ix + y A (x, y)
HK (x, y) = ~vF = (x, y) (2.3.4)
ix y 0 B (x, y)

By applying the Hamiltonian twice and performing the differentiation in x, one gets

2 A =~2 vF2 (kx2 y2 )A (2.3.5)


2
 B =~2 vF2 (kx2 y2 )B . (2.3.6)

The same equations hold for A and B , because the phases eikx x cancel. The general
solution for Eq. (2.3.5) is
v
2
u
i (y) = A ey + B ey , i {A, B}, =
u
tk 2
x . (2.3.7)
(~vF )2

One can derive the solution for sublattice A and then calculate the solution for
sublattice B by using Eq. (2.3.4): B = (kx ) A . In order to determine the
values of A, B, the possible values of and, therefore, the energy eigenvalues , one
can use the boundary conditions and, eventually, normalize the wave function. If

22
2.3. Graphene Nanoribbons

one chooses a unit cell and labels the atomic positions as it is shown in Fig. 2.4, the
boundary conditions can be expressed as

A (0) = 0 (2.3.8)
B (L) = 0. (2.3.9)

Both, A (0) and B (L) refer to the atoms, which are absent due to the geometrical
termination. Eq. (2.3.8) and (2.3.9) yield A = B and
kx
= e2L . (2.3.10)
+ kx
Eq. (2.3.10) supports real solutions only if kx > 0. Because the solutions for K 0
can be obtained by replacing kx kx0 , the condition is
+ kx0
= e2L , (2.3.11)
kx0
which supports solutions only if kx0 < 0. Speaking of the k [, ] instead of
the separation from K or K 0 , real solutions for exist if k > 2
3
or k < 2
3
. Pure
imaginary exist for all k [, ]. They belong to confined waves and not localized
edge states.

Electronic structure investigation using a tight binding model


In the simplest tight binging model, only nearest-neighbor hopping is considered.
Hence, the Hamiltonian is given by:
W
N X
am,n (bm,n + bm1,n + bm,n1 ) + h.c.
X
H = t (2.3.12)
n=1 m=1

where n denotes the graphene unit cell across the width of the ribbon and M along
the periodic x-direction. The translational invariance along the x-axis guarantees
that the Hamiltonian can be diagonalized by the Fourier transformation
1 X ikm
cm,n = e ck,n , c {a, b}. (2.3.13)
M k
M 0
ei(kk )m = kk0 and bm1,n = eika bm,n , one
P
After a short calculation and using
m=1
gets
N X
t X    
H= ak,n bk,n 1 + eika + bk,n1 + h.c. (2.3.14)
M n=1 k
N X
t X
=: Hk . (2.3.15)
M n=1 k

23
2. Theory

Figure 2.8: Illustration of the condition for a state to be the eigen-


function to the eigenvalue zero. Reproduced from [32].

The analytical solutions and more details can be found in [31]. As we are interested
in the zero energy surface states especially, we derive the wave function for this state
within this tight binding model. The second edge is neglected and the following
results are only valid in the limit of an semi-infinite graphene plane. The deviations
should be small for rather large widths. Again, the outermost edge atoms belong to
sublattice B w.l.g.
The wave function on two successive atoms along the periodic direction must
differ only by a phase eik . To fulfill the condition H =  = 0, it is required that
the sum of the amplitudes over the three nearest-neighbors is zero [32]. Considering
the six positions shown in Fig. 2.8, one gets the equations:
eikn + eik(n1) + x =0 (2.3.16)
ikn ik(n+1)
e +e + y =0 (2.3.17)
x + y + z =0. (2.3.18)
x,y and z denote the relative amplitude to the outermost states. Solving these
equations is straightforward [32]:
!!
ik(n+ 21 ) k
x=e 2 cos (2.3.19)
2
!!
1 k
y = eik(n 2 ) 2 cos (2.3.20)
2
!!2
ikn k
z=e 2 cos (2.3.21)
2
Hence, the charge density relative to the

outermost
2m
atoms for the atoms of sublattice
k
B on the zigzag chain m is given by 2 cos( 2 ) . In order to be a physical wave

24
2.3. Graphene Nanoribbons

function, which can be normalized, | cos( k2 )| 12 must be fulfilled. Therefore, |k|


must take values in the interval [ 2
3
, ]. The decay into the bulk is exponential and
the decay length k depends on k through k = ln(4 cos12 (k/2)) . For k and
k 2 3
k approaches 0 and , respectively, meaning that the state is completely
localized on the edge atom in the former case and extends further into the bulk, when
k decreases. This behavior is the same as predicted by the Dirac-Weyl equation in
the previous section.

Conclusion
An insight into the interesting and remarkable feature of graphene nanoribbons with
zigzag edges was given using both, the Dirac-Weyl formalism and a tight binding
approach. Zigzag nanoribbon have two zero energy bands. The corresponding states
are localized on each edge and only live on the sublattice which belongs to the
outermost atom. They decay exponentially into the bulk with a decay length being
a function of k.

25
2. Theory

2.3.2 Armchair Termination


In contrast to zigzag nanoribbons armchair nanoribbons do not have localized edge
states, but their band gap is a function of the width. Armchair nanoribbons can
be categorized in three groups with their band gap obeying 3p+1 > 3p > 3p+2 ,
where p is an integer and 3p, 3p + 1 and 3p + 2 equals the number of dimer lines and
characterizes the width of the nanoribbon. Fig. 2.5 on page 20 shows the unit cell of
an armchair nanoribbon and illustrates the width labeling.

Electronic structure investigation using the Dirac-Weyl equation


This part is also reproduced from [25]. Up to Eq. (2.3.7) the calculations are also
valid for armchair terminated nanoribbons. But as the edge geometry is changed,
we have to assume different conditions. One possibility is to require that the wave
function vanishes where the atoms are absent due to the geometrical termination.
Therefore, the total wave function at this position, which is the sum of the K and
the K 0 valley, has to be zero and the boundary conditions are:

C (0) = C (0) + 0C (0) = 0 (2.3.22)


C (M + 1) = C (M + 1) + C0 (M + 1) = 0, C = A, B. (2.3.23)

Using the general solution from before:

A+B+C +D =0 (2.3.24)
i 4
 
AeL + BeL + e 3a
L
CeL + DeL = 0. (2.3.25)

The additional phase appears because kx and kx0 with |kx | = |kx0 | are measured as
distances from K and K 0 , respectively, and differ by 4
3a
. The boundary conditions
are satisfied with A = D and B = C = 0. Eq. (2.3.25) becomes:
4
e2L = ei 3a L . (2.3.26)

L denotes the sites where the atoms are absent due to geometrical termination and
can be expressed with the number of dimer lines Z by L = Z+1
2
a. The short calcu-
lation
Z +1 4i Z + 1
2 a = 2ni + a, n N
2 3a 2
=3m+j
2i z }| {
(Z + 1) a = 2ni + (Z + 1), m N, j = 0, 1
3
2i j
 
= ikn = n+ , n = n + m
(Z + 1)a 3
shows that possible values of kn are a non-trivial function of the width of the ribbon.
Armchair nanoribbons can be categorized into three different subclasses depending

26
2.4. Bilayer Graphene Nanoribbons

on if the number of dimer lines can be writtenq as 3m, 3m + 1 or 3m + 2. The


energy of a state associated with kn is  = ~vF kn2 + ky2 . States of subclass 3m + 2
(equivalent to j = 0) have a degenerate zero energy solution for n = 0, ky = 0 while
no zero energy solution is possible in the other subclasses.

Conclusion
Armchair nanoribbons do not have localized states like zigzag nanoribbons. But their
band gap varies with the number of dimer lines Z which can be used to characterize
the width. They can be divided into three families where Z is equal to 3p, 3p + 1
and 3p + 2 with p N, respectively. Their band gaps satisfy the inequality
3p+2 < 3p < 3p+1 .

2.4 Bilayer Graphene Nanoribbons


Zigzag Termination
Tight binding calculations have shown [33] that besides monolayer edge states, there
are bilayer and even trilayer states, depending on the particular geometry.
Consecutive layers do not end at the same site but form a step due to the AB
stacking and the fact that the edge atoms belong to the same sublattice. Therefore,
one edge possesses a more "free-standing" character while the other is "on top" of
a ribbon or layer. The "free standing" edge state behaves like the edge state of a
monolayer ribbon discussed before. In contrast, the inner edge is more complex.
The inner edge atom can be part of a dimer or off-dimer site. These cases are
often referred to as - and -edge, respectively, and this nomenclature shall be used
in this work as well. In tight binding calculations [31] clear differences between
- and -edges appear. While -edges show the same properties as the monolayer
edges, -edges have a finite amplitude on both layers and the bulk penetration on
the second layer is larger than on the layer belonging to the edge.
In [33] it was shown that these states also appear at a graphite step which was
modelled by the transition of monolayer to bilayer graphene. In Sec. 3.3.1, common-
alities and differences between the tight binding calculations and the DFT results
are discussed.

Armchair Termination
Besides zigzag bilayer ribbons, armchair bilayer ribbons have also been investigated.
Because edges always consist of two atoms, one of each sublattice, there is only one
type of edge. In [34] the dependence of the band gap on layer distance and ribbon
width are discussed. But, as there are no specific electronic features like the localised
states at zigzag edges, these results cannot explain the properties of a monolayer to
bilayer transition which is discussed in Sec. 3.3.2.

27
2. Theory

28
3. Density Functional Theory Cal-
culations for Graphene systems

This section is organized as follows: Firstly, the results of the graphene monolayer
and the graphene bilayer are briefly discussed. Secondly, graphene nanoribbons with
zigzag and armchair edges are investigated in detail in Sec. 3.2 as they build a
fundamental part of the monolayer-bilayer junctions which are addressed in Sec. 3.3.

All results discussed in this section were obtained by using the Vienna Ab Initio
Simulation Package (VASP) density functional theory code [1214] within the local
density approximation (LDA) [35]. Projector augmented-wave (PAW) potentials [20]
were used as potentials. The correlation energy was parametrized by Perdew and
Zunger [18] using results from Ceperly and Alder [16, 17].

While the k-point sampling is given in each section separately, the energy cutoff
was always set to 520 eV. When specified, the structures were allowed to relax
until the forces on each atom were smaller than 0.02 eV/A. During relaxations the
electronic convergence criterion was set to 104 eV. For all other calculations this
criterion was set to 105 eV.

In all plots and discussions the Fermi energy is equal to zero.

3.1 Extended Graphene

In this section, the basic properties of the graphene monolayer and the graphene
bilayer are reproduced with density functional theory calculations and summarized
as they build the basis of the structures which are discussed later in this work.

29
3. Density Functional Theory Calculations for Graphene systems

Figure 3.1: The honeycomb structure of the graphene monolayer.


a1 and a2 are the two in-plane lattice vectors of graphene. a3 and
the z-direction are perpendicular to the graphene plane and not
shown here. The primitive unit cell is framed by a1 , a2 and the
dotted line. It contains two carbon atoms labeled 1 and 2.

3.1.1 Graphene Monolayer


The lattice vectors were chosen to be
!
3 1
a1 = a , ,0 (3.1.1)
2 2
!
3 1
a2 = a , ,0 (3.1.2)
2 2
!
a3 = a 0, 0, 6 (3.1.3)

with a = 2.42 A being the graphene lattice constant. It is smaller than the experi-
mental lattice constant (aexp = 2.46 A) which is a common error whenever LDA is
used. The two carbon atoms per unit cell were positioned at
!
p1 = 0, 0, 0 (3.1.4)
!
1 1 1
p2 = a1 + a2 = a , 0, 0 . (3.1.5)
3 3 3
During the calculations, the distance between two graphene planes, which occur
due to the fact that VASP [1214] periodically repeats the given unit cell in all
dimensions, was set to more than 14 A. Greater distances up to 20 A did not change
the total energy of the system. The k-point sampling in the self-consistent calculation
was 21 21 1. Because the given structure provides an exact honeycomb lattice
and the lattice constant is known to be accurate for the used PAW potentials [11,20],
the relaxation stopped after the first electronic convergence.

30
3.1. Extended Graphene

10 2
1.5
5
1
Energy/eV

0 EF 0.5
0
-5
-0.5
-10 -1
-1.5
-15
-2

-20
K
reciprocal space
K M
reciprocal space

Figure 3.2: Band structure of the graphene monolayer. The blue


rectangle (left) denotes the vicinity of K at the Fermi energy which
is enlarged on the right-hand side. It clearly shows the linear dis-
persion.

The band structure was calculated at 100 k-points between along each line be-
tween two points of high symmetry. Fig. 3.2 shows the complete band structure and
an enlarged view of the vicinity of K.
From the slope of the bands in Fig. 3.2, the Fermi velocity can be calculated and
turns out to be vf = 1.04 106 ms which is in good agreement with the experimental
value [36]. As already discussed in Sec. 2.2.1, the linear behavior of the bands near
the Fermi energy and the circular symmetry of the Dirac cones leads to a linear
increasing density of states to both, positive and negative energies, while being zero
at the Fermi energy. Fig. 3.3 shows the density of states for all energy values and in
the interval [1 eV, 1 eV].

31
3. Density Functional Theory Calculations for Graphene systems

2.5

Density of states/(1/unit cell)


0.16
Density of states/(1/unit cell)

2 0.14
0.12
1.5 0.1
0.08
0.06
1
0.04
0.02
0.5
0
-1 -0.5 0 0.5 1
0 Energy/eV
-20 -15 -10 -5 0 5 10
Energy/eV

Figure 3.3: Density of states (DOS) of the graphene monolayer. The


blue rectangle (left) denotes the part of the plot which is enlarged
on the right-hand side. In the enlarged plot, one can clearly see
that the DOS is zero at the Fermi energy EF = 0 and its slope is
linear to both sides of EF .

3.1.2 Graphene Bilayer


The unit cell is similar to the one of the monolayer. The lattice vectors are
!
3 1
a1 = a , ,0 (3.1.6)
2 2
!
3 1
a2 = a , ,0 (3.1.7)
2 2
a3 = a (0, 0, 8), (3.1.8)

with a = 2.42 A being the lattice constant. Only the third vector was extended to
keep the distance between two successive layers in the z-direction above 14 A. The
positions of the carbon atoms are
 
r A1 = a 0, 0, 0 (3.1.9)
!
1
r B1 = a , 0, 0 (3.1.10)
3
!
1
r A2 = a , 0, 1.36 (3.1.11)
3
2
 
r B2 = a , 0, 1.36 . (3.1.12)
3

32
3.1. Extended Graphene

Figure 3.4: Two graphene layers in AB stacking: a1 and a2 are


the two in-plane lattice vectors. The primitive unit cell which is
framed by the lattice vectors and the dotted line contains four car-
bon atoms. The atoms are labeled A1, B1 A2 and B2 where A and
B refer to the sublattice and 1 and 2 to the top (black) and bottom
(grey) layer, respectively.

with respect to the lattice vectors. The distance between the two layers is 3.3 A.
Again, the structure is such that the relaxation stops immediately after the first
electronic convergence. In contrast to the monolayer case, the bands corresponding
to the -states do not show a linear but parabolic behavior. This is demonstrated
in Fig. 3.5.
The parabolic dispersion means that the electrons behave like massive particles
in bilayer graphene, while acting like massless particles in monolayer graphene. If the
band structure was rotationally symmetric around K, the parabolic bands would lead
to a constant density of states. But as Fig. 3.6 illustrates, the lines of equal energy
do not form concentric circles around K. Hence, the density of states is not constant.
The jumps at 0.4 eV mark the energies where the two split bands contribute to the
density of states.

33
3. Density Functional Theory Calculations for Graphene systems

10 2
1.5
5
1
Energy/eV

0 EF 0.5
0
-5 -0.5
-1
-10
-1.5

-15 -2
K
-20 reciprocal space
K M
reciprocal space

Figure 3.5: Band structure (left) of bilayer graphene. The blue


rectangle denotes the vicinity of K which is enlarged on the right-
hand side. In contrast to the graphene monolayer (cf. Fig. 3.2), the
dispersion is parabolic for small energies.

0.69

0.55
0.68 0.5
0.45
0.4
0.35
0.67 0.3
k2/ ( /a)

K 0.25
0.2
0.66 0.15
0.1
0.05
0.65

0.64
0.31 0.32 0.33 0.34 0.35 0.36
k1/ ( /a)

Figure 3.6: The plot shows lines of equal energy in the vicinity of
the K-point of the graphene bilayer. There is no strict rotational
symmetry.

34
3.1. Extended Graphene

Density of states/(1/unit cell)


0.3
Density of states/(1/unit cell)

5
0.25
4
0.2

3 0.15

0.1
2
0.05

1 0
-1 -0.5 0 0.5 1
0 Energy/eV
-20 -15 -10 -5 0 5 10
Energy/eV

Figure 3.7: Density of states of the graphene bilayer for all occupied
energies (left) and an enlarged view (right) of an interval around the
Fermi energy, denoted by the blue rectangle in the left-hand plot.
The two jumps at about 0.4 eV mark the energies of the second
valence and conducting band (cf. Fig. 3.5), respectively.

35
3. Density Functional Theory Calculations for Graphene systems

36
3.2. Constrained Graphene

Figure 3.8: A zigzag graphene nanoribbon with hydrogen passiva-


tion: Its unit cell is framed by the rectangle. Its width is Z = 10
zigzag chains which is equal to 19.8 A (neglecting the hydrogen
atoms). The carbon-carbon distance a is about 1.42A. The z-
direction points into the paper plane.

3.2 Constrained Graphene


In the following, graphene nanoribbons with zigzag and armchair edges are inves-
tigated as these structures build the fundamental basis of the monolayer-bilayer
junctions. For both edge configurations, hydrogen passivated and non-passivated
edges are considered in the ideal and the relaxed geometry.
Passivated zigzag graphene nanoribbons feature two flat -bands at the Fermi
level which correspond to localized edge states. These states decay exponentially
into the bulk. Non-passivated zigzag graphene nanoribbons additionally have two -
bands near the Fermi level which originate from the dangling bonds. The interaction
between the -states and the localized -states causes the latter to be dispersive and
lie at higher energies.
Armchair graphene nanoribbons with hydrogen passivation show a direct band
gap which depends crucially on the width of the nanoribbon. The dangling bonds of
the non-passivated armchair nanoribbon form an additional bond between the two
carbon edge atoms. Both systems do not show localized edge states.

3.2.1 Zigzag Termination


Passivated
In this section the electronic properties of zigzag nanoribbons with hydrogen-passivated
edges are discussed. Calculations have been made for both, the ideal geometry and
the relaxed one until the forces were smaller than 0.02 eV/A. By this, the influence
of the relaxation on the localized state can be determined. The k-point sampling in
the self-consistent calculations was 21 3 3. Although the system is not periodic

37
3. Density Functional Theory Calculations for Graphene systems

in the y- and z-direction, three k-points were sampled due to the used Brillouin zone
integration method (tetraedron method with Blchl corrections). We begin with the
ideal geometry.
The lattice vectors are given by:
 
a1 = a 3, 0, 0 (3.2.1)
a2 = a (0, L, 0) (3.2.2)
a3 = a (0, 0, 12) (3.2.3)

with a = 1.413 A the equilibrium carbon-carbon distance for the used PAW poten-
tials [11, 20]. The reason for using the carbon-carbon distance instead of the lattice
constant of graphene is the y-coordinate of the atomic positions are only integers
or half integers which is more convenient. The value of L depends on the width of
the nanoribbon and can be calculated from the number of carbon atoms N or the
number of zigzag lines Z:
3 3
L = 2 + (N/2 2) + 12 = 2 + (Z 2) + 12. (3.2.4)
2 2
The addition of 12 creates the vacuum space to separate the nanoribbons in neigh-
boring cells.

2
1.5
1
Energy/eV

0.5
0 EF
-0.5
-1
-1.5
-2
- /d K K' /d
reciprocal space

Figure 3.9: Band structure of a zigzag graphene nanoribbon with


Z=42 zigzag chains in ideal geometry and with hydrogenated edges.
Note that the two flat bands on the left- and right-hand side of
the plot show almost no dispersion. K (K 0 ) lies at 2 ( 2 ) where
3d 3d
d = 3a 2.42 A is the graphene lattice constant.

38
3.2. Constrained Graphene

0.5
Energy/eV
Z=42
0 Z=16 EF
Z=10

-0.5

-1
- /d K
reciprocal space

Figure 3.10: Vicinity of the K-point of the band structure of three


zigzag graphene nanoribbons with different widths: The splitting of
the flat bands occurs closer to the K-point for larger widths (mea-
sured in the number of zigzag chains Z) because the bulk penetration
of an eigenvalue depends on its value of k and it becomes larger the
closer k is to K.

Fig. 3.9 shows a typical band structure. The almost flat bands in the k intervals
[ d , 2
3d
2
and [ 3d , d ] (d = 3a) correspond to the localized edge states. The point,
]
where the two degenerate bands split, approaches K = 2 3d
as the width of the
nanoribbon is increased. Because the edge states extend further into the bulk the
smaller the difference between k and K becomes, the k-point for which the two states
from both side overlap eventually, lies closer to K for wider nanoribbons. As soon
as the states overlap the degeneracy is lifted. In Fig. 3.10 the vicinity of K is shown
for three different widths. The described effect can be seen clearly.
Fig. 3.11 illustrates the different penetrations into the bulk. The insets show the
charge density (y) = e dx dz |k (x, y, z)|2 of the localized -bands for different
R

k-points. Additionally, (z) is plotted to verify that the edge states have indeed the
pz -orbit shape. The picture in Fig. 3.12 demonstrates that the edge state is almost
completely localized on one sublattice as predicted by the Dirac-Weyl equation in
Sec. 2.3.1.
By projecting the corresponding wave functions onto the atomic orbitals, one can
evaluate the local density of states (LDOS).
( 0 ) |hnlm |0 i|2 , n = 1, .., N
X
dlm (n, ) = (3.2.5)
0

Here, 0 is the energy of the state, |nlm i the lm-orbital of atom n, and |0 i a state

39
3. Density Functional Theory Calculations for Graphene systems

2
b
1.5 (z) (y)

z
0.5
Energy/eV

y
c

0 a
b EF
c

Ribbon
-0.5
a c
(z) (y) (z) (y)
-1
z

z
-1.5
y y
-2
- /d K K' /d
reciprocal space

Figure 3.11: Band structure of a zigzag nanoribbon (Z = 16): The


insets a,b,c show the charge density of the denoted eigenvalues.
All eigenvalues are a linear combinations of pz -orbitals. The bulk
penetration length of the eigenvalues shown in inset a is small
compared tobulk penetration length of those shown in inset b (
and denote the position of the two edges). The two eigenvalues
shown in inset c are delocalized over the entire ribbon.

Figure 3.12: Charge density of an edge state (k = 5 9d


). A and
B mark the two sublattices. The edge state is (almost completely)
localized on one sublattice (B) as it is predicted by the Dirac-Weyl
equation (cf. Sec. 2.3.1, p. 19).

40
3.2. Constrained Graphene

5
LDOS

2
20
1 15
10
0 5
0

-5
-10 Energy/meV
-15
y -20

Figure 3.13: LDOS of a zigzag graphene nanoribbon (Z = 16) with


ideal geometry. It has a U-like shape with one peak above and one
below the Fermi energy EF = 0.

with the energy 0 . As the localized state has an finite amplitude only on one of
the two atoms per zigzag chain, the following quantity was used to guarantee a
consistency throughout this work and create more understandable plots:

Dlm (j, ) = dlm (2j 1, ) + dlm (2j, ), j = 1, .., Z. (3.2.6)

The LDOS which results from Dpz for a small interval around the Fermi energy
is shown in Fig. 3.13. and denote the position of the edges. It has a U-like shape
with peaks at 7 meV and 12 meV and, therefore, it shows that the edge states are
metallic. The states with higher energy belong to values of k which are closer to K
and they show indeed a larger penetration into the bulk.
As can be seen in Fig. 3.14, the decay of a single states is indeed exponential
as predicted by both, the Dirac-Weyl equation and tight-binding models. In exper-
iments the local density of states can be measured by scanning tunnel microscopy.
Edge states which decay into the bulk have been reported [28, 30]. Theoretically,
as one state for each value of k decays exponentially with a different penetration
length, one should see a sum of exponential functions in experiments. But so far the
measurement accuracy is not sufficient. Fig. 3.15 shows both, the DFT data and
sum of exponential functions as derived in the tight-binding part of Sec. 2.3.1.

41
3. Density Functional Theory Calculations for Graphene systems

||2/a.u. k=0.38*2 /a k=0.37*2/a k=0.36*2/a

6 8 10 12 14 6 8 10 12 14 6 8 10 12 14
y/a

Figure 3.14: ||2 of localized states of a zigzag graphene nanoribbon


(Z = 16) for three different values of k. The outermost edge atom
is located at y = 6a. The penetration length increases from left to
right and it is approximately exponential excluding the first atom.

10
tight binding
ab initio

1
DOS

0.1

0.01
6 8 10 12 14 16 18 20
y/a

Figure 3.15: The plot shows the density of states at the Fermi energy
EF = 0. Blue dots represent our ab initio data and the red line is the
analytic solution of a tight binding model from [31]. The agreement
is remarkably good except for the outermost edge atom which is
located at y = 6a.

42
3.2. Constrained Graphene

1.5

1
0.2

0.5
Energy/eV

0.15

0.1

0.05

0 0

-0.05
EF
-0.1

-0.5
-0.15

-0.2
K

-1

-1.5

-2
- /d K K' /d

Figure 3.16: Band structure of a zigzag graphene nanoribbon with


width Z = 16: ideal (red) and relaxed geometry (blue). The differ-
ence is very small as was the change in position.

In the following, the relaxed geometry and differences to the ideal geometry are
discussed. The change in atomic positions is small. Mainly, the atom next to the edge
atom is shifted to the latter as is the hydrogen atom. These changes are of the order
of 102 A. This causes a strain that moves the following atoms towards the edges but
the change in position is very small ( 103 A). Fig. 3.16 shows the band structure
of the relaxed and the ideal nanoribbon for the width Z = 16. Except for a small
shift to lower energies which is reasonable as the relaxed geometry is energetically
more favorable, the band structure does not change. The LDOS (Fig. 3.17) however
changes: Its overall position is shifted to lower energies and the U-like shape is
flatter with two peaks at 10 meV and 7 meV. Between the peaks, the LDOS is
almost constant and, relative to the peaks, much higher than in the ideal case.

43
3. Density Functional Theory Calculations for Graphene systems

3.5

2.5
LDOS

1.5

1
20
0.5 15
10
0 5
0

-5
-10 Energy/meV
-15
y -20

Figure 3.17: LDOS of a zigzag nanoribbon (Z = 16) with relaxed


geometry: The edge states spread over a larger energy interval and
is more constant than in the ideal case (cf. Fig. 3.13, p. 41).

k=0.38*2 /a k=0.37*2/a k=0.36*2/a


||2/a.u.

6 8 10 12 14 6 8 10 12 14 6 8 10 12 14
y/a

Figure 3.18: ||2 of localized states of a zigzag graphene nanoribbon


for three different values of k. The penetration length increases
from left to right. It is always larger compared to the ideal case (cf.
Fig. 3.14, p. 42).

44
3.2. Constrained Graphene

Non-passivated
In the following zigzag nanoribbons without hydrogenated edges are discussed. Due
to geometrical termination, a dangling bond exists on each edge. These sp2 -states
have a crucial influence on the localized -bands which have been discussed in the last
section. In nature, non-passivated nanoribbons are very unstable and their existence
is unlikely even at low temperatures. But the presence of -bonds may be important
in the context of spin-orbit interaction [26]. As before the ideal geometry is addressed
first:

2
c
1.5 (z) (y)

1
z

0.5 b

Energy/eV


0 c EF
b
a
Ribbon
-0.5
a b
(z) (y) (z) (y)
-1
z

-1.5
y y
-2
- /d K K' /d
reciprocal space

Figure 3.19: Band structure of a zigzag graphene nanoribbon


(Z=16) without hydrogen passivation in ideal geometry: The insets
a,b,c show the charge density corresponding to the marked eigen-
values.The pink single peak corresponds the charge density of the
two -bands and proves their localization in the plane of the ribbon.
They are located at the edges and their bulk penetration increases
slightly when approaching . The green charge density corresponds
to the -bands which are shifted upwards in energy compared to
the passivated system (cf. Fig. 3.11, p. 40).

The lattice vectors and atomic positions are the same as in the previous section
except for the removed hydrogen atoms. The band structure is shown in Fig. 3.19. In
contrast to the passivated case two additional bands near the Fermi energy appear.
They are degenerate in the entire Brillouin zone. The charge densities associated
with states in these bands are plotted in the insets (red). Both bands have the

45
3. Density Functional Theory Calculations for Graphene systems

same charge density on both edges. They are symmetric and antisymmetric linear
combinations of states which belong to one edge only. In the z-direction they reach
their maximum in the plane of the nanoribbon. Hence, the states have a sp2 -like
shape and are the dangling bonds that result from the missing passivation. Like the
-states they become less localized by decreasing k, but do not mix or overlap with
the state on the other edge. More interesting is the effect of the dangling bonds on
the localized -bands of which the charge density is also given in the insets (blue):
At k = d they are shifted upwards in energy by almost 0.8 eV. The more delocalized
the states become by decreasing k, the smaller the energy upshift is but still being
more than 0.3 eV at the point where the -bands split. Reasonably, the influence of
the dangling bond is smaller for -states, which are less localized at this position.
Besides that the properties of the -bands do not change compared to the passivated
case. They are still degenerate near the edge of the Brillouin zone and split when the
overlap is significant. Fig. 3.20 shows the LDOS which corresponds to the -bands.
The less the states energy is the less localized it becomes. In contrast to that the
LDOS which corresponds to the -bands, shown in Fig. 3.21, has a U-like shape and
is concentrated on the first two zigzag chains at each edge.

10

8
LDOS

2 0.8
0.7
0.6
0
0.5

0.4
Energy/eV
0.3
y 0.2

Figure 3.20: LDOS of the -bands of a non-passivated zigzag


graphene nanoribbon with ideal geometry: The interaction with
the dangling bonds causes a huge difference compared to the pas-
sivated system (Fig. 3.13). Instead of being concentrated at E = 0,
the LDOS has its maximum at about 0.75 eV.

46
3.2. Constrained Graphene

2.5

2
LDOS

1.5

0.2
0.5
0.1
0
0
-0.1

-0.2
Energy/eV
-0.3
y -0.4

Figure 3.21: LDOS of the -bands of a non-passivated zigzag


graphene nanoribbon with ideal geometry. It was obtained by the
projection on the 2s-orbital which is orthogonal to the pz -orbital
and, therefore, does not include contributions from -states.

Compared to the passivated nanoribbons, the shifts in position due to the re-
laxation are larger. The outermost atom is shifted inwards while the next atom is
slightly moved outwards (both by 102 A). The effect of the relaxation on the
band structure is shown in Fig. 3.22.
Both, the - and the -bands lower their energy significantly, the -bands by ca.
0.2 eV and the -bands by about 0.35 eV. Hence, the stronger coupling between the
pz -orbitals due to the shift in position is able to reduce the effect of the dangling
bonds on their energy. The charge densities do not change besides the slightly
changed position of the outermost edge atom.

Conclusion
In this section the properties of zigzag nanoribbons were discussed. The passivated
nanoribbon features localized states on both edges which correspond to two flat
bands in the region 2
3d
< |k| < d . Their appearance and properties can be described
by a tight-binding model or the Dirac-Weyl equation (see Sec. 2.3.1). The relaxation
causes only minor changes.
The non-passivated nanoribbon has a sp2 dangling bond at each edge which

47
3. Density Functional Theory Calculations for Graphene systems

2
1.5
1
Energy/eV

0.5

0 EF
-0.5

-1
-1.5
-2
- /d K K' /d
reciprocal space

Figure 3.22: Band structure of the ideal (red) and relaxed (blue)
zigzag graphene nanoribbon (Z=16) without hydrogen passivation.
All bands lower their energy significantly but the energy of the two
-bands is still higher than the Fermi energy EF = 0.

interacts with the localized -state such that the latter is shifted to higher energies.
This upshift is strongly reduced by the relaxation.

48
3.2. Constrained Graphene

Figure 3.23: An armchair graphene nanoribbon with hydrogen pas-


sivation: Its unit cell is framed by the rectangle. The number of
atoms is much larger compared to zigzag nanoribbons for identical
widths (cf. Fig. 3.8, p. 37).

3.2.2 Armchair Termination


In this section the properties of nanoribbons with armchair edges is discussed. The
ideal and the relaxed geometry are considered with and without hydrogen passiva-
tion. The lattice vectors are given by:

a1 = a (3, 0, 0) (3.2.7)
a2 = a (0, L, 0) (3.2.8)
a3 = a (0, 0, 12). (3.2.9)

Like before, a = 1.413 A is the carbon-carbon distance for the used PAW potentials
[11, 20] and L can be determined by

3
L= (Z 1) + 12 = 3(N 2) + 12, (3.2.10)
2
N being the number of carbon atoms and Z the number of dimer lines. The addition
of 12 creates the vacuum space between the nanoribbons in two successive cells.
Armchair nanoribbons do not feature localized edge states, but show an interesting
behavior of the band gap as already discussed in Sec. 2.3.2. The k-point sampling
was 15 3 3. Firstly, the passivated case is addressed:

Passivated
Fig. 3.24 shows the band structure for an armchair nanoribbon of width Z = 29,
30 and 31. These respectively correspond to width W = 34.3 A, 35.5 A and 36.7 A.
Although the width is increased, the band gap widens.
The band gap as a function of the number of the dimer lines Z is shown in
Fig. 3.25, the corresponding values are also given in Tab. 3.1. In contrast to the

49
3. Density Functional Theory Calculations for Graphene systems

Z=29 Z=30 Z=31


2
1.5
1
Energy/eV 0.5
0 EF
-0.5
-1
-1.5
-2

Figure 3.24: Band structure of three different armchair graphene


nanoribbons: The band gap is a function of the width Z (number of
dimer lines). It is smallest for the 3p + 2-family (3 9 + 2 = 29) and
largest for the 3p + 1-family (31). This hierarchy is predicted by
the Dirac-Weyl equation (Sec. 2.3.1, p. 19), but the ab initio results
show that a band gap exists for all three families.

1
3p+1
3p
3p+2
0.8
band gap/eV

0.6

0.4

0.2

0
10 15 20 25 30 35 40 45 50
Number of dimer lines

Figure 3.25: Band gaps at the Dirac point for armchair graphene
nanoribbons with ideal geometry of various widths: The band gap of
each of the three families decreases as 1/Z with Z being the number
of dimer lines across the armchair nanoribbon which is indicated by
the colored lines.

50
3.2. Constrained Graphene

Z 17 18 19 29 30 31 43 44 45
W/A 19.6 20.8 22.0 34.3 35.5 36.7 51.4 52.6 53.8
i /eV 0.122 0.362 0.594 0.074 0.226 0.369 0.268 0.050 0.153
r /eV 0.175 0.385 0.640 0.104 0.240 0.392 0.284 0.061 0.160
(r i )/eV 0.053 0.023 0.046 0.030 0.014 0.023 0.016 0.011 0.007

Table 3.1: Band gaps of armchair graphene nanoribbons for different


widths with ideal (i ) and relaxed (r ) geometry.

results from Sec. 2.3.2 even in the case Z = 3p + 2, p N the band gap does
not vanish. This was also shown in tight-binding calculations which included third-
nearest-neighbor hopping [24]. All three families show a 1/Z decay of the band
gap.
Another reason for the opening of a band gap is edge distortion resulting from
relaxation [24]. The changes in position due to relaxation are again very small: the
two edge atoms are shifted towards the nanoribbon and each other by 102 A. The
effect on the inner atoms is 103 A or less. Fig. 3.26 and Tab. 3.1 show the band gaps
for relaxed armchair nanoribbons with the same widths. Comparing these results
with the unrelaxed, one sees: All band gaps are increased by relaxation but the
effect becomes weaker the wider the nanoribbon is. The increase is smallest for the
3p-family and largest for the 3p + 2-family.

51
3. Density Functional Theory Calculations for Graphene systems

1
3p+1
3p
3p+2
0.8
band gap/eV

0.6

0.4

0.2

0
10 15 20 25 30 35 40 45 50
Number of dimer lines

Figure 3.26: Band gaps of armchair graphene nanoribbons at the


Dirac point for relaxed geometry: The band gap of each of the three
families decreases again as 1/Z (Z number of dimer lines across the
armchair nanoribbon) as indicated by the lines. All gaps are larger
compared to the ideal geometry (cf. Fig. 3.25, Tab. 3.1).

Z=29 Z=30 Z=31


2
1.5
1
Energy/eV

0.5
0 EF
-0.5
-1

-1.5
-2

Figure 3.27: Band structures of three different non-passivated arm-


chair graphene nanoribbons with ideal geometry. Additional -
bands appear near the Fermi energy EF = 0 and the bands which
build the band gap in the passivated case (cf. Fig. 3.24) are pushed
upwards in energy by ca. 0.4 eV.

52
3.2. Constrained Graphene

Non-passivated
In the following, the non-passivated armchair nanoribbons are discussed. As for
zigzag nanoribbons -bands appear which result from the dangling bonds. Fig. 3.27
shows the band structures for the widths Z = 29, 30 and 31. Because the number of
non-passivated edge atoms is now four, two per edge, four additional bands appear
near the Fermi level. In the band structure they show up as two pairs of different
energy: One lies below the Fermi level, the other above. The edges are equal due
to symmetry. Hence, the degenerate bands are symmetric and antisymmetric linear
combinations of the states on each edge. The two bands with lower energy form a
bond between the two edge atoms, while the other two form an anti-bond. Both are
shown in Fig. 3.28. The Dirac cone is shifted from the Fermi-energy to higher energy.
The opening of the Dirac cone follows the same hierarchy as in the passivated case.
The exact values are given in Tab. 3.2. But here the opening at the Dirac cone does
not coincide with the band gap at the Fermi energy because of the -bands.
For passivated armchair nanoribbons we have seen that the width and small edge
distortion have a crucial influence on the opening of the Dirac cone in the center of
the Brillouin zone. Therefore, we expect a larger influence on the band structure
for non-passivated nanoribbons as the change in geometry compared to the ideal is
larger than for passivated nanoribbons. Fig. 3.29 shows the relaxed structure and
the arrows denote the change in position magnified by 10. The two outermost atoms
are shifted inwards and towards each other like in the passivated case. But here, the
change is much larger and of the order of 101 A.
The effect on the band structure can be seen in Fig. 3.30. The two bands which
correspond to the bond between the edge atoms lowers its energy by ca. 1.5 eV while
the other two bands are lifted by 0.4 eV. Both do not change their general shape.
The Dirac cone is shifted back to the Fermi level, although it does not always build a
direct band gap for small widths because of the -band at the edges of the Brillouin
zone. While in all other cases 29 < 30 < 31 hold for i being the opening of the
Dirac cone, this is not true for non-passivated, relaxed armchair nanoribbons. 31 is

Figure 3.28: The dangling bonds at the edge of a non-passivated


armchair graphene nanoribbon which result from the missing passi-
vation: They form a bond anti-bond pair.

53
3. Density Functional Theory Calculations for Graphene systems

Z 17 18 19 29 30 31
W/A 19.6 20.8 22.0 34.3 35.5 36.7
i /eV 0.071 0.415 0.545 0.042 0.253 0.338
r /eV 0.362 0.218 0.636 0.196 0.136 0.406
(r i )/eV 0.291 -0.197 0.091 0.154 -0.117 0.068

Table 3.2: -band gaps of non-passivated armchair graphene


nanoribbons at the -point for different widths with ideal (i ) and
relaxed (r )geometry.

Figure 3.29: Relaxed structure of an armchair graphene nanoribbon


(Z = 19): The arrows show the change in position relative to the
ideal geometry enlarged by a factor of 10.

Z=29 Z=30 Z=31


2
1.5
1
Energy/eV

0.5

0 EF
-0.5
-1
-1.5
-2

Figure 3.30: Band structures of three different non-passivated arm-


chair graphene nanoribbons with relaxed geometry: The hierar-
chy of the band gaps changed from 3p+2 < 3p < 3p+1 (cf.
Fig. 3.24,3.27) to 3p < 3p+2 < 3p+1 .

54
3.2. Constrained Graphene

still the largest but 29 is increased by a factor of 5 compared to the ideal geometry.
Simultaneously, 30 is decreased by a factor of 2. The exact values for these and
other widths are given in Tab. 3.2.

Conclusion
The band gap of a passivated armchair nanoribbon depends on the width which is
often expressed by the number of dimer lines Z. It obeys 3p+2 < 3p < 3p+1 with
i being the band gap. In contrast to the simple approach given in Sec. 2.3.2, a
finite band gap is present for all Z, but the ranking is correct. Within the families
the band gap falls as 1/Z with increasing Z.
The two sp2 dangling bonds of non-passivated armchair nanoribbons form a bond
- anti-bond pair. Due to the relatively strong change in geometry, the ranking of the
band openings at is changed to 03p < 03p+2 < 03p+1 . Here, 0i is not necessarily
the band gap because of the - bands which result from the dangling bonds.

55
3. Density Functional Theory Calculations for Graphene systems

56
3.3. MonolayerBilayer Graphene junctions

Figure 3.31: Graphene monolayerbilayer junction with zigzag ter-


mination: The rectangle denotes the unit cell. For simplicity, the
system shown here is much smaller than the one discussed below.
The z-direction points into the paper plane.

3.3 MonolayerBilayer Graphene junctions


A nanoribbon is positioned on top of a graphene layer to model the transition from
monolayer graphene to bilayer graphene. In all cases the ideal geometry was assumed
to be the ideal ribbon on an ideal graphene layer in AB stacking. The equilibrium
distance of bilayer graphene was chosen as distance between the nanoribbon and the
graphene layer. As in the previous part the two basic edge geometries, zigzag and
armchair, are discussed and the effect of passivation and relaxation are investigated.
As in free standing zigzag graphene nanoribbons (Sec. 3.2.1), the terminated
graphene layer features localized -bands. The two possible edge configurations
with zigzag termination (cf. Fig. 3.31) interact differently with the graphene layer
below and induce different states in the latter. While the energy difference between
the two edge configurations is very small in the passivated case, it is much larger in
the non-passivated one where the energy difference between the two configurations
is more than 0.1 eV.
The armchair terminated monolayer-bilayer-monolayer junction shows a combi-
nation of the graphene monolayer and the graphene bilayer spectrum. In the contin-
uous graphene layer, monolayer states exists which do not interact with the armchair
nanoribbon on top. The band curvature and the exact electronic structure depends
on the width of the bilayer region.

3.3.1 Junction with Zigzag Termination


Due to the AB stacking, two interfaces are possible: The edge atom can be on top of
an atom in the graphene layer (dimer site) or on top of honeycomb hole (off-dimer

57
3. Density Functional Theory Calculations for Graphene systems

site). The former will be referred to as -edge, the latter as -edge.


The lattice vectors are similar to those of the free zigzag nanoribbon:

a1 = a ( 3, 0, 0) (3.3.1)
a2 = a (0, L, 0) (3.3.2)
a3 = a (0, 0, 14) (3.3.3)

with a = 1.413 A the carbon-carbon distance. While a1 was not changed and a3
trivially increased in length to keep the distance between two successive layers in
z-direction, L has to take into account the continuous graphene layer. The structure
which has to be repeated is basically an element of an armchair nanoribbon unit cell.
Hence, to achieve a continuous layer by the periodic repetition of VASP [1214] , L
must be a multiple of 3 (The length of the unit cell of an armchair nanoribbon in the
periodic direction is also 3a). Further, L depends on the width of the monolayer and
the bilayer areas one wants to consider. Because the left and the right edge atom of
the nanoribbon belong to different sublattices, they automatically form an - and a
-edge. The k-point sampling in the self-consistent calculation is 21 5 3.
It is known from tight-binding models of bilayer ribbons [31] and monolayer-
bilayer interfaces [8, 33] that the existence of an edge state has a crucial influence
on the underlying graphene layer. A localized state with a wider penetration into
the bulk is induced in the latter. The -edge shows no such state in tight-binding
calculations. In the following the band structure, the local density of states (LDOS),
and the different edge states will be discussed.

Passivated
Fig. 3.32 shows the band structure of a zigzag nanoribbon of width W = 23a (Z = 16)
on a graphene layer with ideal geometry. The monolayer region has a width of 16a.
The inset gives an enhanced view of the vicinity of K. As in the case of a free
nanoribbon flat bands in the regions 2 3d
|k| d exist which correspond to the
edge states. But instead of being almost completely flat they form a eye-like shape
near K and K 0 because of their admixture with the graphene bulk states.
In Fig. 3.33 the band structure is shown again and the partial charge density
is given for three k-points in insets. Each partial charge density belongs to one of
the two localized bands. The insets a and b show that each of the bands belongs
to one edge only. This is in contrast to the free standing ribbon were the two edges
where completely equivalent which is not the case here due to the two different edge
configurations. Moreover, in all insets the charge density is distributed differently
in the z-direction, although the difference is not as huge as one would expect from
the tight binding results, where only the -edge state has a finite amplitude on both
layers. Hence, the edge state located at the -site also couples to the graphene layer.
As already discussed in Sec. 2.3.1 and 3.2.1 the bulk penetration becomes larger for
k approaching K, as well as the amplitude on the graphene layer. Both effects can

58
3.3. MonolayerBilayer Graphene junctions

1.5

1
0.3

0.5
Energy/eV

0.2

0.1

0 0
EF
-0.1

-0.2

-0.5 -0.3
K

-1

-1.5

-2
- /d K K' /d

Figure 3.32: Band structure of a zigzag graphene nanoribbon (Z =


16) on a graphene layer. It is similar to the band structure of the
free standing zigzag nanoribbon (cf. Fig. 3.9, p. 38). But the two
flat bands at the Fermi energy build an eye-like shape near K and
K 0.

59
3. Density Functional Theory Calculations for Graphene systems

1
b
(z) (y)

0.5

z
Energy/eV

b c
0 EF
a
a b
Ribbon
c
Layer

a c
(z) (y) (z) 5*(y)
-0.5
z

z
y y
-1
- /d K K' /d
reciprocal space

Figure 3.33: Band structure of a zigzag graphene nanoribbon (Z =


16) on a graphene layer (ideal): The insets a,b,c show the partial
charge densities at the denoted k-points. The blue colored states
belong to the lower of the two bands. In contrast to the free standing
zigzag nanoribbon (cf. Fig. 3.11 on p. 40), each band belongs to
one edge because the edges are no longer equal by symmetry. For
k K, more amplitude is located on the graphene layer and the
penetration length is increased.

be seen by comparing the insets a and b. For k = K, which is shown in inset c,


the state is a bulk state without localization and more than half of its amplitude is
located on the graphene layer.
Fig. 3.34 shows the LDOS of the nanoribbon near the Fermi energy. One can see
that the edges differ: The -edge shows a slightly lower LDOS amplitude but because
its also wider in energy, the integrated LDOS in the shown energy interval is larger
at the -edge. For both edges, the LDOS has a U-like shape and the penetration into
the bulk is larger for higher energies. This can be explained by the band structure:
The flat bands have a slight slope upwards towards K and K 0 , respectively, and
the states closer to K and K 0 are less localized. Like for the free standing zigzag
nanoribbon the states of both edges are metallic. The slope is smaller for |k| d
and |k| 23d
which explains the U-like shape.
In Fig. 3.35 the LDOS of the graphene layer is shown. Note that the scale on
the z-axis is smaller by a factor of more than 20. The presence of an edge induces

60
3.3. MonolayerBilayer Graphene junctions

60

50

40
LDOS

30

20

20
10
15
10
0 5
mono 0

-5
bi
-10 Energy/meV
-15
y mono
-20

Figure 3.34: LDOS of a passivated zigzag graphene nanoribbon


(Z = 16) on a graphene layer with ideal geometry. The LDOS
at the -edge is wider and slightly lower in energy.

61
3. Density Functional Theory Calculations for Graphene systems

2.5

2
LDOS

1.5

0.5 20
15
10
0 5
mono 0

-5
bi
-10 Energy/meV
-15
y mono
-20

Figure 3.35: LDOS of the graphene layer on which a passivated


zigzag graphene nanoribbon (Z = 16) is placed. The geometry is
ideal. The two edge geometries induce different states. At the -
edge the state has a larger bulk penetration and it is located inside
the bilayer region only. The state at the -edge is more localized,
but extends in the monolayer and the bilayer region. Compared to
the LDOS of the nanoribbon (Fig. 3.34), the amplitude is lower by
a factor of 20.

62
3.3. MonolayerBilayer Graphene junctions

Figure 3.36: Charge density of an -edge state: It extends over


both layers in the bilayer region and its amplitude is located only
on the sublattice of the outermost edge atoms.

a LDOS in the underlying layer regardless of whether it is an or type. But


the properties of the induced state differ: While the state on the -site has a finite
amplitude only on one sublattice (the same which the outermost edge atoms belongs
to, see Fig. 3.36) in the bilayer region, the -site state extends over both sublattices
near the honeycomb hole (see Fig. 3.37). The penetration length is larger on the
side. It extends up to six lattice constants for high energies. The maximum for higher
energies is located one graphene cell further in the bulk compared to the low energy
maximum which is in agreement with the prediction from tight binding results [33].
On the side, the state consists of three peaks of which one is in the monolayer
region and two in the bilayer region. For some energies, the latter are merged. This
possibly happens due to admixture with bulk states which are also visible in both,
mono- and bilayer regions.
The changes in geometry caused by the relaxation are similar to those of the
free-standing zigzag nanoribbon. The hydrogen atom and the carbon atom next to
the edge atom are moved towards the latter by 102 A while the bulk and the
graphene layer atoms are offset in various positions by 103 A or less.
Fig. 3.38 shows the band structure of both, the ideal and the relaxed system.
The differences are relatively small: Only the slope of the localized bands increases
significantly. This can also be seen in Fig. 3.40: The U-like shape is smoothed out
over a larger energy range and the LDOS is almost constant between the two peaks.
The -edge state is lower in energy by 10 meV. Of course, this broadening also
happens on the graphene layer, shown in Fig. 3.41. The bulk penetration is similar
to the ideal structure but becomes harder to tell exactly because the amplitudes are
smaller due to the smoothening and the bulk states are of comparable amplitude.
Tight binding calculations and experiments have demonstrated that the LDOS
of the two edges differ. While the TB results show that it is larger at the -edge

63
3. Density Functional Theory Calculations for Graphene systems

Figure 3.37: Charge density of a -edge state: It is located in the


monolayer and the bilayer region and it has a finite amplitude on
both sublattices of the continuous graphene layer.

1.5

1
0.3

0.5
Energy/eV

0.2

0.1

0 0
EF
-0.1

-0.2

-0.5 -0.3
K

-1

-1.5

-2
- /d K K' /d

Figure 3.38: Band structure of the ideal (red) and relaxed (blue)
zigzag graphene nanoribbon on a graphene. The inset gives an en-
hanced view of the vicinity of the K-point. As for the free standing
zigzag nanoribbon (Fig. 3.16, p. 43), the changes in the band struc-
ture due to the relaxation are very small, as are the changes in the
atom positions.

64
3.3. MonolayerBilayer Graphene junctions

1
b
(z) (y)

0.5
z
Energy/eV

b c
0 a
a b
Ribbon
EF
c
Layer

a c
(z) (y) (z) 5*(y)
-0.5
z

y y
-1
- /d K K' /d
reciprocal space

Figure 3.39: Band structure of a zigzag graphene nanoribbon (Z =


16) on a graphene layer (relaxed): The insets a,b,c show the partial
charge densities at the denoted k-points. The blue states belongs
to the lower band. In inset b, one can see that the interaction of
the nanoribbon with the graphene layer is different for the two edge
configurations because the state at the -edge has more amplitude
on the graphene layer.

65
3. Density Functional Theory Calculations for Graphene systems

35

30

25
LDOS

20

15

10
30
5
20
10
0
mono 0

bi -10
Energy/meV
-20
y mono
-30

Figure 3.40: LDOS of a passivated zigzag graphene nanoribbon


(Z = 16) on a graphene layer with relaxed geometry. As in the
free-standing case (Fig. 3.17, p. 44), the edge states spread over a
larger energy interval. The two edges differ by ca. 10 meV.

66
3.3. MonolayerBilayer Graphene junctions

1.6

1.4

1.2

1
LDOS

0.8

0.6

0.4
30
0.2 20
10
0
mono 0

bi -10
Energy/meV
-20
y mono
-30

Figure 3.41: LDOS of the graphene layer on which a passivated


zigzag graphene nanoribbon (Z = 16) is placed. The two edges differ
in energy as well as in the spatial distribution which is qualitatively
equal to the ideal case (Fig. 3.41). But as they extend over a larger
energy interval, their amplitude becomes comparable to the bulk
states which are clearly visible in the monolayer and the bilayer
region.

67
3. Density Functional Theory Calculations for Graphene systems

[33], experimental data leads to the opposite assumption [30]. Within our ab initio
calculations the LDOS is larger at the -edge, but the difference is less than 2%
relative to the maximum value and, therefore, too small to be responsible for the
large difference that has been observed experimentally. Since small geometric shifts
can have a huge influence on the amplitude and the energetic position of the edge
states as the comparison of ideal and relaxed geometry shows, a possible explanation
of the different results from experiments is that a geometric disturbance due to the
small length of the zigzag edge caused the high LDOS.

Non-passivated
The non-passivated system shows a couple of differences compared to the passivated.
The ideal geometry is the same as before except for the hydrogen atoms. The band
structure is shown in Fig. 3.42.

2
c
1.5 (z) (y)

1
z

0.5
Energy/eV

y
b


0 c
EF
b
a Ribbon
-0.5 Layer
a b
(z) (y) (z) (y)
-1
z

-1.5
y y
-2
- /d K K' /d
reciprocal space

Figure 3.42: Band structure of non-passivated zigzag graphene


nanoribbon on a graphene layer with ideal geometry. The insets
a,b,c show the charge density for the particular states. Blue and
red correspond to the lower two bands which belong to in-plane
states. Green and purple to the upper two bands which belong to
the localized -states.

Due to the interaction with the sp2 -dangling bonds, shown in Fig. 3.43, the two
-bands (of which two states are also shown in Fig. 3.43) are pushed up in energy. In
contrast to the free-standing ribbon this effect is not of equal strength on both sides:

68
3.3. MonolayerBilayer Graphene junctions

(a) Dangling bond at the edge (b) Dangling bond at the edge

(c) state at the edge (d) state at the edge

Figure 3.43: Charge densities of the dangling bonds (a,b) and the
localized states (c,d) at the - and -edge. While the charge
density of the -state at the -edge extends over both layers, the
-state at the -edge does not.

Even though the influence of the underlying layer on the -bands was very small in
the passivated case, it is very strong here. Both bands extends over an interval of
0.5 eV compared to 15 meV in the passivated case. Instead of a U-like shape, the
LDOS, shown in Fig. 3.44, has maxima at the highest energies, which is 0.8 eV for
the and 0.65 eV for the -edge, and decays to smaller energies. Again, the band
structure explains this behavior: The modulus of slope of the bands is small near
the BZ edge and increases when k approaches K. The fact that the dangling bond
lies on top of a hexagon hole is the only difference that explains the huge difference
in energy between the two edges, since the underlying graphene layer did not cause
such a strong difference in energy. This is even more remarkable in connection with
Fig. 3.45. It shows the projection of the states associated with the -bands on the
2s-orbital. Since these dangling bond states are 2sp2 -hybrid orbitals, this projection
allows to evaluate the LDOS which belongs to them without contributions of the
-bands. The location in energy does not differ and the band structure also shows
that neither the graphene layer nor the interaction with the -bands influences the
dangling bonds.
During the relaxation, both edges are bended towards the graphene layer. The
bending is stronger on the -side by a factor of 4. The change in position is shown
in Fig. 3.46.
Fig. 3.47 shows the band structure of the relaxed geometry and the charge den-
sity for certain states. The general behavior is similar to that of the ideal structure.

69
3. Density Functional Theory Calculations for Graphene systems

30

25

20
LDOS

15

10

5 0.8
0.7
0 0.6
mono
0.5
bi 0.4 Energy/eV
0.3
y mono
0.2

Figure 3.44: -LDOS of the non-passivated zigzag graphene


nanoribbon (Z = 16) on a graphene layer. As can also be seen
in Fig. 3.42 the two edges differ in energy. The peak of the LDOS
at the -edge lies about 0.1 eV below the peak at the -edge.

70
3.3. MonolayerBilayer Graphene junctions

16

14

12

10
LDOS

4
0.1
2 0.05
0
0 -0.05
mono -0.1

-0.15
bi
-0.2 Energy/eV
-0.25
y mono
-0.3

Figure 3.45: -LDOS of the non-passivated zigzag graphene


nanoribbon (Z = 16) on a graphene layer. While the -bands be-
have differently for the - and -edge (cf. Fig. 3.44), the dangling
bonds do not.

(a) -edge (b) -edge

Figure 3.46: Geometrical change caused by the relaxation in the


system consisting of an non-passivated zigzag graphene nanoribbon
on a graphene layer: The arrows show the change made enlarged by
15. The shifts are much larger at the -edge.

71
3. Density Functional Theory Calculations for Graphene systems

2
c
1.5 (z) (y)

1
z

0.5
Energy/eV

a
y
b
c
0 EF
b
a
Ribbon
-0.5 Layer
a b
(z) (y) (z) (y)
-1
z

-1.5
y y
-2
- /d K K' /d
reciprocal space

Figure 3.47: Band structure of a non-passivated zigzag graphene


nanoribbon on a graphene layer with relaxed geometry. The insets
a,b,c show the charge density for the particular states. The - and
the -bands which belong to the two edges, respectively, differ in
energy due to the geometrical difference. Red denotes the lower
-band, blue the upper. Green corresponds to the lower -bands
which belongs to the -edge and purple to the upper -band which
belongs to the -edge.

72
3.3. MonolayerBilayer Graphene junctions

30

25

20
LDOS

15

10

0.6
5
0.5
0.4
0 0.3
mono 0.2

0.1
bi
0 Energy/eV
-0.1
y mono
-0.2

Figure 3.48: -LDOS of the non-passivated zigzag graphene


nanoribbon (Z = 16) on a graphene layer with relaxed geometry.
The energetic difference between the - and the -edge is larger
than in the ideal case (Fig. 3.44). The large bulk penetration can
be seen clearly for values of k close to K which correspond to the
low energy states shown here.

Because of the change in geometry, the charge density of the -edge state is shifted
further towards the layer than that of the -edge. The -bands differ by less than
0.1eV in their energy. The -bands lowered their energy by about 0.3 eV and reach
their maxima at 0.25 eV () and 0.45 eV () above the Fermi energy. The energy dif-
ference between the two bands is slightly increased, the extent of 0.5 eV is unchanged.
The LDOS shown in Fig. 3.48 also demonstrates the difference in energy between
the two edges. At low energies (where k K), one can see the large bulk penetra-
tion. Fig. 3.49 shows the LDOS. Despite the huge differences of the -states, the
-states differ hardly.

73
3. Density Functional Theory Calculations for Graphene systems

12

10

8
LDOS

0.1
2
0
-0.1
0
mono -0.2

bi -0.3
Energy/eV
-0.4
y mono
-0.5

Figure 3.49: -LDOS of the non-passivated zigzag nanoribbon


(Z = 16) on a graphene layer with relaxed geometry. The dif-
ference between the two edges is very small compared to the huge
difference in the -states (Fig. 3.48).

74
3.3. MonolayerBilayer Graphene junctions

Conclusion
The graphene monolayerbilayer junction with zigzag termination offers two possible
interface configurations which are referred to as - and -edge. Besides the localized
edge state, the -edge features a localized state on the continuous graphene layer in
the bilayer region. Its penetration length is larger than that of the edge state but its
amplitude is much smaller. At the -edge a state is also induced on the graphene
layer. It is localized in the monolayer and the bilayer region and is caused by the
interaction of the edge state which lies on top of the center of a hexagon with the
pz -orbitals of the same hexagon.
The band structure is similar to the one of the free standing zigzag nanoribbon
except for an eye-shaped dispersion near K and K. It does not show monolayer states
with a linear dispersion.
As in case of the non-passivated zigzag nanoribbon, the edge states are shifted
upwards in energy but the shift is much lower for the -edge than for the -edge.
The relaxation causes larger shifts in position at both edges than in the passivated
case.

75
3. Density Functional Theory Calculations for Graphene systems

76
3.3. MonolayerBilayer Graphene junctions

Figure 3.50: Monolayerbilayer junction with armchair termination:


The unit cell is denoted by the rectangle. The z-direction points into
the paper plane.

3.3.2 Junction with Armchair Termination


Because an armchair termination involves both atoms of sublattice A and B, there
is only one edge type and the two edges which appear in the investigated system are
equivalent. The lattice vectors are given by:
a1 = a (3, 0, 0) (3.3.4)
a2 = a (0, L, 0) (3.3.5)
a3 = a (0, 0, 14) (3.3.6)

with a = 1.413 A being the carbon-carbon distance. L has to be chosen such that
a continuous
graphene layer can be constructed. Therefore, L must be a multiple
of 3. a1 and a3 were not changed except for an enhancement of a3 to keep the
vacuum space large enough. The k-point sampling in the self-consistent calculations
is 15 5 3. All band structure plots only show the inner third of the Brillouin
zone.

Passivated
Since VASP is only able to treat periodic systems, the systems which are discussed
in the following can be seen as a monolayerbilayermonolayer interfaces. Such
interfaces are investigated [7, 37] because of the assumption to be used as switches
by applying a gate voltage in the bilayer region. The relaxation causes only very
small changes. Therefore, the relaxed structure is discussed.
The free-standing armchair nanoribbon showed a band gap which obeyed the
relation 3p+2 < 3p < 3p+1 . Comparing this with the missing band gap of
graphene, one can think of the 3p + 2-family behaving more like graphene than the
other two families.

77
3. Density Functional Theory Calculations for Graphene systems

Band structure Charge densities at


0.4 Monolayer state Ribbon state Hybrid state ribbon
graphene
0.3

0.2

z
0.1
Energy/eV

(z) (z) (z)


0 Monolayer state

-0.1

(y)
-0.2

-0.3

-0.4
y
reciprocal space

Figure 3.51: Band structure (left) of a Z=29 armchair graphene


nanoribbon on a graphene layer: The behaviour of the lines was
obtained by the partial charge densities and are shown to guide the
eye. The band structure shows both, a monolayer and a bilayer
spectrum.
Charge densities (right) of the four bands near the Fermi energy at
.

Z = 29 : Fig. 3.51 shows the band structure of a nanoribbon with Z = 29 dimer


lines on a graphene layer. Four bands are present near the Fermi level: two with a
linear dispersion touching at 15 meV. The other two have a parabolic shape and
are separated having their maximum and minimum at 10meV and 20meV, respec-
tively. To assign the eigenvalues to the different bands, their partial charge density
was used. On the right hand side, Fig. 3.51 shows the charge densities of the four
corresponding eigenvalues at . The lowest eigenvalue has its whole amplitude on the
monolayer. The second and third show a similar charge density distribution on the
nanoribbon and the graphene layer. The highest is mostly located on the nanoribbon.

Z = 30 : The band structure, shown in Fig. 3.52, also has two parabolic and two
linear bands but the parabolic bands are less flat than for Z = 29. While the linear
bands touch at 10meV, the parabolic bands have their minimum and maximum
at 0 and 10meV, respectively. The charge densities prove that the linear bands are
localized on the graphene layer while the parabolic bands have their amplitude on
both, the nanoribbon and the graphene layer. Therefore, the band structure is a
combination of the monolayer and the bilayer spectrum and the states will be called
monolayer and bilayer states in the following.
Z = 31 : The band structure is given in Fig. 3.53. The bands near the Fermi level
can still be separated into two linear and two non-linear bands. Unfortunately, the k-
point sampling could not be increased sufficiently to determine the band shape more
precisely. The linear bands touch 10meV below the Fermi energy and are located
on the graphene layer. The other two which are spread over the nanoribbon and

78
3.3. MonolayerBilayer Graphene junctions

Band structure Charge densities at


0.4 Monolayer state Bilayer state ribbon
graphene
0.3

0.2

z
0.1
Energy/eV

(z) (z)
0 Monolayer state

-0.1

(y)
-0.2

-0.3

-0.4
y
reciprocal space

Figure 3.52: Band structure (left) of a Z=30 armchair graphene


nanoribbon on a graphene layer: It shows a monolayer and a bi-
layer spectrum. The behaviour of the lines was obtained by the
partial charge densities and are shown to guide the eye. The charge
densities (right) show that two states have their amplitudes on both
layers (bilayer states) and two are only located on the continuous
graphene layer (monolayer states). The shaded area denotes the
bilayer region. The monolayer states are not influenced by the pres-
ence of the armchair nanoribbon.

the graphene layer have their maximum and minimum at 0 and 17meV, respectively.
Because of their charge distribution , the bands will also be called monolayer and
bilayer states.
Like the free-standing armchair nanoribbon, the monolayerbilayermonolayer
interface composed by a nanoribbon on a graphene layer shows a different behavior
for the widths Z = 3p, 3p + 1 and 3p + 2 with p N.
It is remarkable that even in presence of an armchair nanoribbon the monolayer
state on the graphene layer is not disturbed. The wave functions in the armchair
nanoribbons depend on the width as the model in Sec. 2.3.2 predicts. Putting the
nanoribbon on top of the graphene layer leads to an interaction between the states
on the nanoribbon and the graphene layer. The way and strength of this interaction
depends on the spatial distribution and the overlap between the states. Hence, the
interaction and the resulting band structure is different for the three families of
nanoribbons when put on a graphene layer which can be seen by comparing the
three band structures shown above. The free-standing nanoribbon with Z=29 has
the smallest band gap and its states resemble the monolayer states more than those
of the nanoribbons with widths Z=30 and 31. Therefore, the coupling between the
3p + 2 family and the graphene layer is closest to the graphene bilayer and the
resulting bilayer bands show the strongest curvature while the 3p + 1 family shows
the opposite behavior. But in all cases, the interaction leads to the combination
of the monolayer and the bilayer spectrum. Probably, this has consequences for

79
3. Density Functional Theory Calculations for Graphene systems

Band structure Charge densities at


0.2 Monolayer state Bilayer state ribbon
graphene
0.15

0.1

z
0.05
Energy/eV

(z) (z)
0 Monolayer state

-0.05

(y)
-0.1

-0.15

-0.2
y
reciprocal space

Figure 3.53: Band structure (left) of a Z=31 armchair graphene


nanoribbon on a graphene layer: The behaviour of the lines was
obtained by the partial charge densities and are shown to guide the
eye.
Charge densities for the four bands near the Fermi energy at
(right). The shaded area denotes the bilayer region. Although the
band structure does not show a combination of the monolayer and
the bilayer spectrum as for the smaller widths, the states can be
characterized as in Fig. 3.52.

transmission coefficients and possible applications in electronic devices. It is very


likely that the continuous monolayer state causes a high transmission coefficient
through the bilayer region. If and how these states are influenced by a perpendicular
gate voltage cannot be predicted by the investigations shown in this section.

Non-passivated
The differences between relaxed and ideal geometry are similar to the differences
already discussed in Sec. 3.2.2. There is no coupling between the dangling bonds
and the graphene layer as in the zigzag case. Therefore again, only the relaxed
structure is discussed in the following.
The band structures for Z = 29 (Fig. 3.54) and 30 (Fig. 3.55) show a combination
of the monolayer and the bilayer spectrum: There are four bands near the Fermi
energy of which two are located on the graphene layer and the other two on both
layers. The curvature of the bilayer bands is almost identical. A comparison with
the results from Sec. 3.2.2 (Tab. 3.2, p. 54) shows that the band gaps of these two
nanoribbons are also very similar. This supports the assumption that the curvature
is correlated with the band gap of the free standing nanoribbon as already states
above. In contrast to the passivated system, the maximum and minimum of the
bilayer states are lower in energy than the Dirac point of the two monolayer bands.
The Dirac point is located above the Fermi level. Due to the edge distortion and
the missing passivation, the bilayer states are energetically more favorable than the

80
3.3. MonolayerBilayer Graphene junctions

Band structure Charge densities at


0.4 Monolayer state Bilayer state ribbon
graphene
0.3

0.2

z
0.1
Energy/eV

(z) (z)
0 Monolayer state

-0.1

(y)
-0.2

-0.3

-0.4
y
reciprocal space

Figure 3.54: Band structure of a Z=29 non-passivated armchair


graphene nanoribbon on a graphene layer: The behaviour of the
lines was obtained by the partial charge densities and are shown
to guide the eye. As in the passivated case, the band structure
is a combination of the monolayer and the bilayer spectrum. But
here, the monolayer bands touch above the minimum of the bilayer
valence band.

Band structure Charge densities at


0.4 Monolayer state Bilayer state ribbon
graphene
0.3

0.2
z

0.1
Energy/eV

(z) (z)
0 Monolayer state

-0.1
(y)

-0.2

-0.3

-0.4
y
reciprocal space

Figure 3.55: Band structure of a Z=30 non-passivated armchair


graphene nanoribbon on a graphene layer: The behaviour of the
lines was obtained by the partial charge densities and are shown
to guide the eye. The band structure is also a combination of the
monolayer and the bilayer spectrum.

81
3. Density Functional Theory Calculations for Graphene systems

Band structure Charge densities at


0.4 ribbon
graphene
0.3

0.2

z
0.1
Energy/eV

(z) (z)
0

-0.1

(y)
-0.2

-0.3

-0.4
y
reciprocal space

Figure 3.56: Band structure of a Z=31 non-passivated armchair


graphene nanoribbon on a graphene layer: It does not show a com-
bination of the monolayer and the bilayer spectrum as the band
structure of the smaller widths do. The states of all bands are lo-
cated on both layers in the monolayer and the bilayer region as the
charge density plots on the right-hand side show.

monolayer states.
Fig. 3.56 shows the band structure for Z = 31. Here, a band gap exists and
no characterization into monolayer and bilayer states is possible. The two valance
and conduction bands are almost identical in both, dispersion and the corresponding
charge density distribution. On the right-hand side of Fig. 3.56, the following charge
densities are shown: (z) of the two valance bands which are identically distributed,
(z) of the two conduction bands and (y) for one of each types to demonstrate that
all bands are affected by the presence of the nanoribbon. The corresponding free
standing ribbon (cf. Tab. 3.2, p. 54) shows a larger band gap than the nanoribbons
with the widths Z = 29 and 30. Additionally, the edge distortion leads to an spatial
width such that the coupling between the layers does not allow the existence of
monolayer states.

Conclusion:
The monolayer-bilayer-monolayer junction with armchair edges shows two different
types of states near the Fermi level. Two bands show the linear dispersion of free
standing graphene monolayer. Their states are located on the continuous graphene
layer in the monolayer and the bilayer regions and they are not disturbed by the
presence of the nanoribbon. The other two bands correspond to states which are
located on the nanoribbon and the graphene layer and show a parabolic dispersion.
The curvature of the dispersion depends on the width of the bilayer region and it is
strongest for the 3p + 2 family in the passivated case. The free standing nanoribbons
of this family show the smallest band gap and closely resemble the electronic structure
of free standing graphene. Hence, it is reasonable that the coupling to the graphene

82
3.3. MonolayerBilayer Graphene junctions

layer resembles the coupling in graphene bilayer mostly for this family.
The non-passivated structures behave similarly. The structures with bilayer
widths Z = 29 and 30 show two monolayer bands and two bilayer bands. The
curvature of the latter is almost identical for the two widths. But relative to the
bilayer bands the monolayer bands are shifted to higher energies such that their
touching point lies above the Fermi level. The Z = 31 case does not show a com-
bination of the monolayer and the bilayer spectrum but a band gap. All for bands
correspond to states which extend over the armchair nanoribbon and the graphene
layer.

83
3. Density Functional Theory Calculations for Graphene systems

84
4. Conclusion

Free standing nanoribbons and nanoribbons placed on graphene have been studied
in this thesis by density functional theory calculations. These calculations were done
with the Vienna Ab Initio Simulation Package (VASP) [1214]. The nanoribbons
placed on graphene were used to describe the graphene monolayerbilayer junction
and monolayer-bilayer-monolayer structure.
Graphene is the two dimensional allotrope of carbon in which the carbon atoms
form a honeycomb lattice. One of the special properties of graphene is the linear
dispersion of the electronic states at the Fermi energy. One dimensional strips of
graphene are called graphene nanoribbons. Their electronic properties depend on
the edge geometry of which the two simplest are the zigzag edge and the armchair
edge.
Zigzag nanoribbons have localized edge states on each edge at the Fermi energy
which correspond to two flat bands in the regions 2 3a
< |k| < a , a being the lattice
constant of graphene. The edge states decay exponentially into the bulk and can
be described very well by both, the Dirac-Weyl equation and tight binding models.
Armchair nanoribbons have a direct band gap at . The size of the band gap allows
to divide armchair nanoribbons into three families depending on their width which
is often measured in the number of dimer lines Z. Z can be represented as 3p, 3p + 1
or 3p + 2 with p N and the band gaps i obey 3p+2 < 3p < 3p+1 .
The monolayerbilayer junction with a zigzag edge shows electronic properties
which go beyond the edge state found in zigzag nanoribbons. If the edge is part of
a dimer site (called -edge), a localized state is induced in the graphene layer in the
bilayer region. Its bulk penetration is much larger than that of the nanoribbon edge
state but its overall amplitude is smaller by a factor of 20. The second possible edge
configuration where the edge atom is on top of a honeycomb hole of the graphene layer
(called -edge) shows a different behavior. A state is also induced in the graphene
layer but it is localized within three lattice constants in both, the monolayer and the
bilayer region. While the -edge state has already been described by tight binding
calculations, the -edge state has not.
If the edges remain non-passivated, the formerly flat bands are shifted upwards
in energy. While the two bands are still degenerate for a free standing ribbon, their
energy differs when the nanoribbon is put on a graphene layer. Then, the -edge
state is energetically more favorable.

85
4. Conclusion

The systems consisting of an armchair nanoribbon on a graphene layer does not


have two different edges. Neither are edge states present which could interact with
the graphene layer. The band structure shows four bands near the Fermi energy,
two with a linear dispersion and two with an approximately parabolic shape. While
the latter correspond to states which have their amplitude on the nanoribbon and
the graphene layer, the linear bands correspond to states which have their amplitude
on the graphene layer only. The presence of the nanoribbon does not even influence
these monolayer state in the bilayer region which may have a huge influence on the
transmission through such a junction. The strength of the parabolic band curvature
depends on the number of dimer lines of the bilayer region and it is strongest for the
3p + 2 family.
The non-passivated case is similar to the passivated except for the four additional
bands near the Fermi level. In one case (Z = 31, 3p + 1 family), the non-passivated
system shows a band gap and the states of all four bands extend over both layers. It
is probably a strong edge distortion which causes a coupling between the graphene
layers that disturbs the monolayer states in the continuous graphene layer and does
not. The corresponding free standing armchair nanoribbon has a band gap which is
much larger than the band gap of the free standing nanoribbons with Z = 29 and 30.

Comparing the two edge geometries, one can see that the coupling between a
nanoribbon and a continuous graphene layer depends on the edge configuration.
The coupling is rather weak for the armchair terminated system. A monolayer state
exists despite the presence of the nanoribbon. However, the coupling is different
for different widths of the bilayer region which is correlated with the modulo 3
behavior of armchair nanoribbons. In zigzag terminated junctions, the coupling is
much stronger due to the localized edge states which spread over both layers. Due
to the strong influence, no monolayer state exists in the bilayer region.
The results described above are not directly transferable to the junction between
epitaxial graphene to quasi free standing bilayer graphene on SiC where the latter
is obtained by local hydrogen intercalation from epitaxial graphene. In this system,
the Fermi level does not coincide with the Dirac points but is located above and
below, respectively.
Further research is needed to include effects of the different Fermi levels in mono-
layer and bilayer region for the system on SiC as described above. The influence
of the existing monolayer state on the transmission probability through armchair
terminated monolayerbilayer junctions is of great interest, especially for the possi-
bility to build electronic devices with graphene, and may also be the focus of further
research.

86
References

[1] Philip Richard Wallace. The band theory of graphite. Physical Review,
71(9):622, 1947.
[2] KSA Novoselov, Andre K Geim, SVb Morozov, Da Jiang, MIc Katsnelson, IVa
Grigorieva, SVb Dubonos, and AAb Firsov. Two-dimensional gas of massless
dirac fermions in graphene. nature, 438(7065):197200, 2005.
[3] Alexander Mattausch and Oleg Pankratov. Ab initio study of graphene on sic.
Phys. Rev. Lett., 99:076802, Aug 2007.
[4] C. Riedl, C. Coletti, T. Iwasaki, A. A. Zakharov, and U. Starke. Quasi-free-
standing epitaxial graphene on sic obtained by hydrogen intercalation. Phys.
Rev. Lett., 103:246804, Dec 2009.
[5] J Ristein, S Mammadov, and Th Seyller. Origin of doping in quasi-free-standing
graphene on silicon carbide. Physical review letters, 108(24):246104, 2012.
[6] Andre K Geim and Konstantin S Novoselov. The rise of graphene. Nature
materials, 6(3):183191, 2007.
[7] J. W. Gonzlez, H. Santos, M. Pacheco, L. Chico, and L. Brey. Electronic
transport through bilayer graphene flakes. Phys. Rev. B, 81:195406, May 2010.
[8] Zi-xiang Hu and Wenxin Ding. Edge states at the interface between monolayer
and bilayer graphene. Physics Letters A, 376(4):610615, 2012.
[9] Pierre Hohenberg and Walter Kohn. Inhomogeneous electron gas. Physical
review, 136(3B):B864, 1964.
[10] Walter Kohn and Lu Jeu Sham. Self-consistent equations including exchange
and correlation effects. Physical Review, 140(4A):A1133, 1965.
[11] Peter E Blchl. Projector augmented-wave method. Physical Review B,
50(24):17953, 1994.
[12] G. Kresse and J. Hafner. Ab initio molecular dynamics for liquid metals. Phys.
Rev. B, 47:558561, Jan 1993.

87
References

[13] G. Kresse and J. Furthmller. Efficiency of ab-initio total energy calculations


for metals and semiconductors using a plane-wave basis set. Computational
Materials Science, 6(1):15 50, 1996.

[14] G. Kresse and J. Furthmller. Efficient iterative schemes for ab initio total-
energy calculations using a plane-wave basis set. Phys. Rev. B, 54:1116911186,
Oct 1996.

[15] R.M. Dreizler and E.K.U. Gross. Density Functional Theory. Springer Verlag,
Berlin, 1990.

[16] D. Ceperley. Ground state of the fermion one-component plasma: A monte carlo
study in two and three dimensions. Phys. Rev. B, 18:31263138, Oct 1978.

[17] D. M. Ceperley and B. J. Alder. Ground state of the electron gas by a stochastic
method. Phys. Rev. Lett., 45:566569, Aug 1980.

[18] J. P. Perdew and Alex Zunger. Self-interaction correction to density-functional


approximations for many-electron systems. Phys. Rev. B, 23:50485079, May
1981.

[19] Randolph Q. Hood, M. Y. Chou, A. J. Williamson, G. Rajagopal, R. J. Needs,


and W. M. C. Foulkes. Quantum monte carlo investigation of exchange and
correlation in silicon. Phys. Rev. Lett., 78:33503353, Apr 1997.

[20] G. Kresse and D. Joubert. From ultrasoft pseudopotentials to the projector


augmented-wave method. Phys. Rev. B, 59:17581775, Jan 1999.

[21] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim.


The electronic properties of graphene. Rev. Mod. Phys., 81:109162, Jan 2009.

[22] Liying Jiao, Li Zhang, Xinran Wang, Georgi Diankov, and Hongjie Dai. Nar-
row graphene nanoribbons from carbon nanotubes. Nature, 458(7240):877880,
2009.

[23] Dmitry V Kosynkin, Amanda L Higginbotham, Alexander Sinitskii, Jay R


Lomeda, Ayrat Dimiev, B Katherine Price, and James M Tour. Longitudi-
nal unzipping of carbon nanotubes to form graphene nanoribbons. Nature,
458(7240):872876, 2009.

[24] D. Gunlycke and C. T. White. Tight-binding energy dispersions of armchair-


edge graphene nanostrips. Phys. Rev. B, 77:115116, Mar 2008.

[25] L. Brey and H. A. Fertig. Electronic states of graphene nanoribbons studied


with the dirac equation. Phys. Rev. B, 73:235411, Jun 2006.

88
References

[26] M. P. Lpez-Sancho and M. C. Muoz. Intrinsic spin-orbit interactions in flat


and curved graphene nanoribbons. Phys. Rev. B, 83:075406, Feb 2011.

[27] D. Porezag, Th. Frauenheim, Th. Khler, G. Seifert, and R. Kaschner. Construc-
tion of tight-binding-like potentials on the basis of density-functional theory:
Application to carbon. Phys. Rev. B, 51:1294712957, May 1995.

[28] Chenggang Tao, Liying Jiao, Oleg V Yazyev, Yen-Chia Chen, Juanjuan Feng,
Xiaowei Zhang, Rodrigo B Capaz, James M Tour, Alex Zettl, Steven G Louie,
et al. Spatially resolving edge states of chiral graphene nanoribbons. Nature
Physics, 7(8):616620, 2011.

[29] Young-Woo Son, Marvin L. Cohen, and Steven G. Louie. Energy gaps in
graphene nanoribbons. Phys. Rev. Lett., 97:216803, Nov 2006.

[30] Yousuke Kobayashi, Ken-ichi Fukui, Toshiaki Enoki, and Koichi Kusakabe.
Edge state on hydrogen-terminated graphite edges investigated by scanning tun-
neling microscopy. Phys. Rev. B, 73:125415, Mar 2006.

[31] Eduardo V. Castro, N. M. R. Peres, J. M. B. Lopes dos Santos, A. H. Castro


Neto, and F. Guinea. Localized states at zigzag edges of bilayer graphene. Phys.
Rev. Lett., 100:026802, Jan 2008.

[32] Kyoko Nakada, Mitsutaka Fujita, Gene Dresselhaus, and Mildred S Dressel-
haus. Edge state in graphene ribbons: Nanometer size effect and edge shape
dependence. Physical Review B, 54(24):17954, 1996.

[33] Eduardo V Castro, NMR Peres, and JMB Lopes dos Santos. Localized states
at zigzag edges of multilayer graphene and graphite steps. EPL (Europhysics
Letters), 84(1):17001, 2008.

[34] Kai-Tak Lam and Gengchiau Liang. An ab initio study on energy gap of
bilayer graphene nanoribbons with armchair edges. Applied Physics Letters,
92(22):223106, 2008.

[35] J. P. Perdew and Alex Zunger. Self-interaction correction to density-functional


approximations for many-electron systems. Phys. Rev. B, 23:50485079, May
1981.

[36] R. S. Deacon, K.-C. Chuang, R. J. Nicholas, K. S. Novoselov, and A. K. Geim.


Cyclotron resonance study of the electron and hole velocity in graphene mono-
layers. Phys. Rev. B, 76:081406, Aug 2007.

[37] Johan Nilsson, A. H. Castro Neto, F. Guinea, and N. M. R. Peres. Transmission


through a biased graphene bilayer barrier. Phys. Rev. B, 76:165416, Oct 2007.

89
References

90
Acknowledgement

I would like to thank Prof. Dr. Pankratov for his advice and support and for giving
me the opportunity to this thesis.

Additionally, I would like to thank Stephan Hensel for his guidance and many
helpful discussions and PD Dr. Bockstedte for helping me with VASP-related
problems and many other questions.
Erklrung
Hiermit versichere ich, dass ich die vorliegende Arbeit selbststndig verfasst und keine
anderen als die angegebenen Quellen und Hilfsmittel benutzt habe, dass alle Stellen
der Arbeit, die wrtlich oder sinngem aus anderen Quellen bernommen wurden,
als solche kenntlich gemacht sind und dass die Arbeit in gleicher oder hnlicher Form
noch keiner Prfungsbehrde vorgelegt wurde.

Erlangen, den 25. Juni 2014 Unterschrift

Você também pode gostar