Master Thesis Optical Properties of Pentacene and Picene: University of The Basque Country WWW - Mscnano.eu
Master Thesis Optical Properties of Pentacene and Picene: University of The Basque Country WWW - Mscnano.eu
eu
Basque Country
MASTER THESIS
Optical properties of
pentacene and picene
by
Jeiran Jokar
Supervisors:
Prof. Angel Rubio Secades
Dr. Matteo Gatti
Dr. Pierluigi Cudazzo
14 September 2011
Acknowledgement
I would like to express my gratitude to all those who gave me the possibility
to complete this thesis. This work would not have been possible without
the support from Prof. Dr. Angel Rubio under whose guidance I chose this
topic. I would like to gratefully acknowledge the supervision of my advisor,
Dr. Matteo Gatti, who has been abundantly helpful and has assisted me in
numerous ways. I specially thank him for his infinite patience. The discus-
sions I had with him were invaluable. I would like to say a big thanks to
Dr. Pierluigi Cudazzo, who helped me a lot.
I would like to thank Donostia International Physics Center and the Uni-
versity of the Basque Country for financial supports.
3
Contents
1 Introduction 7
2 Theory 9
2.1 DFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 The Hohenberg-Kohn theorems . . . . . . . . . . . . . 11
2.1.2 The Kohn-Sham method . . . . . . . . . . . . . . . . . 12
2.1.3 Local Density Approximation . . . . . . . . . . . . . . 13
2.1.4 Pseudopotential approximation . . . . . . . . . . . . . 14
2.2 Time Dependent Density Functional Theory . . . . . . . . . . 15
2.2.1 The Runge-Gross theorem . . . . . . . . . . . . . . . . 15
2.2.2 Time-Dependent Kohn-Sham Theory . . . . . . . . . . 15
2.2.3 xc functionals . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.4 DFT vs TDDFT . . . . . . . . . . . . . . . . . . . . . 16
2.2.5 Linear response theory . . . . . . . . . . . . . . . . . . 17
2.2.6 Casida’s equations . . . . . . . . . . . . . . . . . . . . 20
2.2.7 Electronic excitations by means of time-propagation . 20
3 Results 23
3.1 Ground state properties . . . . . . . . . . . . . . . . . . . . . 23
3.1.1 Convergence study . . . . . . . . . . . . . . . . . . . . 23
3.1.2 Structural optimisation . . . . . . . . . . . . . . . . . 26
3.1.3 Kohn-Sham orbitals and energy levels . . . . . . . . . 28
3.2 Absorption spectra from time propagation . . . . . . . . . . . 33
3.2.1 Convergence study . . . . . . . . . . . . . . . . . . . . 33
3.3 Absorption spectra from Casida equation . . . . . . . . . . . 38
4 Conclusion 45
bibliography 47
List of Figures 49
List of Tables 51
5
Chapter 1
Introduction
Organic molecular solids have attracted considerable interest for both funda-
mental reasons and possible technological applications. In fact they are very
promising materials for a wide range of applications [1], as their properties
can be efficiently tailored for their use in (opto)electronic devices [2] such as
field-effect transistor [3], light-emitting diodes [4][5], photovoltaic cells [6][7],
etc. Among the different classes of organic molecular solids, the most stud-
ied one in the last years is probably that of the aromatic molecules. The
interest for this class of molecular crystals recently increased after the dis-
covery of superconductivity in potassium doped picene and potassium doped
coronene at critical temperature Tc =18 K [8] and 15 K respectively [9].
Since the constituents of molecular solids are molecular units interacting
through weak Van der Waals forces their physical properties, as for example
dynamical, electronic and optical properties, are to a large extent governed
by the individual molecular units [10]. Thus, in principle, by studying the
isolated molecule we have access also to the corresponding condensed states.
In this work we present a first-principles study based on density func-
tional theory of the electronic and optical properties of two aromatic molecules
namely picene and pentacene. This will be a good starting point for the un-
derstanding of the electronic and optical properties of the related molecular
crystals.
Aromatic molecules are obtained by combining benzene rings in differ-
ent ways. In particular picene and pentacene both with molecular formula
C22 H14 consist of five benzene rings in two different configurations. In pen-
tacene the benzene rings are arranged in a linear manner (Fig. 3.7(a)) while
in picene they are arranged in a zigzag manner (Fig. 3.7(b)). Since C atoms
present 4 valence electrons (2 in the 2s state and 2 in the 2p state) they form
4 covalent bonds while H atoms only one bond involving their 1s electron.
Due to the planar structure in picene and pentacene as well as in benzene, for
each carbon atom the s orbital combines with two p (px , py ) orbitals. The
resulting three hybrid orbitals are oriented at 120◦ angles to each other.
7
8 CHAPTER 1. INTRODUCTION
These orbitals are called sp2 hybrid orbitals because they are composed of
one s and two p orbitals. If the sp2 orbitals of different C atoms overlap
in phase between them or with the H s orbital they interact constructively
to form a bonding molecular orbitals. The corresponding electron density
results centred along the line connecting the atomic nuclei. This type of
bond is called a σ bond. When they interact out of phase they give rise to
anti-bonding states called σ ∗ state. The remain pz orbitals (one for each C
atoms) combine among them giving the so called π bond. The π bond is
parallel to the line connecting the C nuclei and the related electron density is
localized above and below the line connecting the nuclei. The corresponding
antibonding states are called π ∗ states. Since the overlap between in-plane
atomic orbitals is larger than that between out of plane atomic orbitals, σ
bonds result stronger than the π ones [11]. Thus the electronic structure of
picene and pentacene can be described in terms of σ states responsible for
in plane C-C and C-H bonds and π states responsible for out of plane C-C
bonds. However being the π bonds the weakest, the chemical and optical
properties of these molecules are set mainly by π and π ∗ states.
The present work is organized as follows: in chapter one we discuss
the theoretical tools used in this thesis, namely DFT and TDDFT while
in the second chapter we present our results. Here, first we address the
structural and electronic properties of picene and pentacene then we focus on
their optical properties evaluating the absorption spectra with two different
method: performing time propagation and solving the Casida equation. The
last chapter is devoted to our conclusions.
Chapter 2
Theory
2.1 DFT
The first goal is to find the ground state of the molecules. In fact, by knowing
the ground state we can obtain the atomic structure of the molecules, and the
electronic properties (the orbital wavefunctions and energies). Moreover, the
ground state of the system will be the starting point to employ perturbation
theory (see sec. (2.2.5)) to study the optical response.
In order to find the ground state of a system of M atoms (with N elec-
trons), one has to solve the Schrödinger equation:
N N N
1 X 2 1 X ∇2I 1 X 1
H=− ∇i − +
2 2 MI 2 |ri − rj |
i=1 I=1 i6=j=1
X ZI 1X ZI ZJ
− + , (2.2)
|ri − RI | 2 | RI − RJ |
iI I6=J
where the index i runs over electrons and I runs over nuclei. These are just
the kinetic terms of nuclei and electrons, plus the instantaneous Coulomb
interaction between all pairs of bodies.
One of the first proposal to solve Eq. (2.1) was published in 1927 by
Max Born and Julius Robert Oppenheimer. They observed that the nuclei
in the system move much slower than the electrons because they are more
massive [12]. From this, they assumed that the electrons are following the
slow motion of the nuclei. For each fixed ionic configuration, they are in
their ground state. The total wavefunction Ψ(r1 , . . . , rN , R1 , . . . , RM ) is
9
10 CHAPTER 2. THEORY
Two years later the russian physicist Vladimir Aleksandrovich Fock mod-
ified the Hartree’s method by allowing the wavefunction to be represented
by a Slater determinant [15]. This approach is commonly known as the
Hartree-Fock method (HF). The trial wavefunction ΨHF is antisymmetric
with respect to the interchange of both spin and space coordinates of any
two electrons.
ψ1 (r1 ) ψ2 (r1 ) · · · ψN (r1 )
1 ψ1 (r2 ) ψ2 (r2 ) · · · ψN (r2 )
ΨHF = √ . .. .. (2.5)
N ! ..
. .
ψ1 (rN ) ψ2 (rN ) · · · ψN (rN )
• The second theorem states that a universal functional for energy E[n]
in terms of the density n(r) can be defined, valid for any external
potential Vext (r). For any particular Vext (r), the exact ground state
energy of the system is the global minimum value of this functional,
and the density n(r) that minimizes the functional is the exact ground
state density n0 (r).
The functional E[n] alone is sufficient to determine the exact ground
state energy and density. In general, excited states of the electrons
must be determined by other means.
12 CHAPTER 2. THEORY
The Hohenberg-Kohn theorem is valid only for the densities that are the
ground state of some potential. These densities are called V -representable.
While the V -representability condition assesses the existence of the one-
to-one mapping between density and Vext , the Hohenberg-Kohn theorem
establishes its uniqueness. In the worst case, the potential has degenerate
ground states such that the given density is representable as a linear com-
bination of the degenerate ground-state densities (the density is then called
ensemble-V -representable) [17].
The Hohenberg-Kohn theorem has been then generalized in many ways:
to cases where the ground states is degenerate, or where the system is spin-
polarized,to relativistic DFT [18], to multicomponent DFT,to DFT for su-
perconductors [19], and so on.
occ
X
n(r) = |Φi (r)|2 . (2.7)
i
n(r0 ) 3 0
Z
VH (r) = d r, (2.10)
|r − r0 |
2.1. DFT 13
is the electron-ion interaction (see Eq. (2.3)). εi and Φi denote the eigen-
values and eigenfunctions of the Kohn-Sham equation, respectively. The
wavefunctions calculated by Eq. (2.8) yield the charge density by Eq.( 2.7).
Hence, the Kohn-Sham equation must be solved self-consistently. It seems
that Eq. (2.8) plays a role of Schrödinger equation of one-electron wave-
function, but the idea underlying these equation is quite different. While
in Hartree, Hartree-Fock, or configuration interaction the key variable is
the many body wavefunction Ψ(r1 , ..., rN ) and the ground-state is found by
solving the many body Hamiltonian (2.3), in DFT the key variable is the
much simpler electronic density n(r) and the ground-state energy is found
by minimizing the energy functional or, equivalently, by solving the Kohn-
Sham equations. However, the Kohn-Sham orbitals and energies do not have
a physical meaning. The Slater determinant obtained using Kohn-Sham or-
bitals is different from the Slater determinant that is the solution of the
Hartree-Fock equations.
where εHEG
xc (n(r)) is the exchange-correlation energy per electron in a ho-
mogeneous electron gas of density n [20]. This equation is, by construction,
exact for the homogeneous electron gas and we can expect it to work well
for systems where the density has small spatial variations, or where the
electron-electron interaction is well-screened.
The domain of applicability of LDA has been unexpectedly found to
go much beyond the nearly-free electron gas and accurate results can be
obtained for inhomogeneous systems like atoms or molecules.
14 CHAPTER 2. THEORY
∇2
∂
i Φi (r, t) = − + VKS [n](r, t) Φi (r, t). (2.16)
∂t 2
16 CHAPTER 2. THEORY
VKS [n](r, t) = Vext (r, t) + VHartree [n](r, t) + Vxc [n](r, t). (2.17)
With respect to Kohn-Sham DFT, all the potentials become here time-
dependent.
The potential Vxc is a functional of the the density n(r, t). This means that
in principle to obtain Vxc at given r̄ and t̄ one needs to know the density at
all the points r and times t < t̄. The exact expression of Vxc as a functional
of the density is unknown. At this point we are obliged to perform an
approximation [21][23][24].
2.2.3 xc functionals
Adiabatic approximations In adiabatic approximations, we neglect the
dependence on the past. The approximate Vxc (r, t) only depend on the
static [n](r) evaluated at n = n(r, t) to obtain
density at time t. We can use Vxc
adiabatic (r, t).
Vxc
By inserting the LDA functional in Vxc adiabatic (r, t) we obtain the so-called
The ALDA assumes that the xc potential at the point r, and time t is equal
to the xc potential of a (static) homogeneous-electron gas (HEG) of density
n(r, t).
mechanical action
Z t1
dthΦ(t)|i∂t − Ĥ(t)|Φ(t)i = A[n]. (2.20)
t0
3. In
PNDFT,KSthe density just depends on the space coordinate n(r) =
2
i=1 |Φi (r)| . In TDDFT, the density depends on the time and
space coordinates n(r, t) = N KS 2
P
i=1 |Φi (r, t)| .
The same difference holds for the potentials. VKS in DFT
n(r0 )
Z
δExc [n]
VKS (r) = Vext (r) + dr0 0
+ , (2.21)
|r − r | δn(r)
while in TDDFT
n(r0 , t) δAxc [n]
Z
VKS (r, t) = Vext (r, t) + dr0 + . (2.22)
|r − r0 | δn(r, t)
HKS (r)ΦKS KS KS
i (r) = εi Φi (r). (2.23)
Note that the response function that enters Eq. (2.26), χKS , is the density
response function of a system of non-interacting electrons and is, conse-
quently, much easier to calculate than the full interacting χ. In terms of the
unperturbed stationary Kohn-Sham orbitals it reads :
∞
X Φj (r)Φ∗j (r0 )Φk (r0 )Φ∗k (r)
χKS (r, r0 , ω) = lim (fk − fj ) , (2.27)
η→0+ ω − (j − k ) + iη
jk
δVKS (r, t) = δVext (r, t) + δVHartree (r, t) + δVxc (r, t). (2.28)
2.2. TIME DEPENDENT DENSITY FUNCTIONAL THEORY 19
δVxc (r, ω)
fxc (r, r0 , ω) = , (2.29)
δn(r0 , ω)
where
HEG d HEG
fxc (n) = V (n) (2.34)
dn xc
is just the derivative of the xc potential of the homogeneous electron gas.
The ALDA kernel is local both in the space and time coordinates [23].
20 CHAPTER 2. THEORY
where
Z Z
0
hq|fHxc (ω)|q i = 3
d r d3 r0 ϕq (r)fHxc (r, r0 , ω)ϕq0 (r0 ). (2.37)
cσ(ω)
S(ω) = . (2.39)
2π 2
We may define the dynamical dipole polarizability αij (ω) as the quotient
of the induced dipole moment in the direction i with the applied external
electrical field in the direction j , which yields:
Z Z
αij = − d3 r d3 r0 xi χ(r, r0 , ω)x0j . (2.43)
δhX̂i i(ω)
Z
1
αij = − =− d3 rxi δn(r, ω). (2.45)
k k
Chapter 3
Results
In principle the quantities that appear in Eq. (3.1) are defined for a contin-
uous set of infinite points. However, we cannot deal with an infinite number
of points on a computer. To implement Eq. (3.1) on a computer we have two
possibilities. We can either expand the functions on a basis and truncate
the expansion in a finite number of basis elements.
Or we can represent the wavefunction on a discrete and finite grid of points.
In both cases we introduce an error which we can control by doing a conver-
gence analysis. Octopus uses the grid representation. So we need to check
that the grid is accurate with respect to its spacing and its extension, and
at the same time we should balance the accuracy with the time that the
calculation takes to run.
So we have to study the convergence of the results with respect to two pa-
rameters: first the grid spacing (i.e. the distance between mesh points) and
second the radius of the simulation box that contains the molecule.
23
24 CHAPTER 3. RESULTS
Energy [eV]
−3646
−3648
−3650
−3652
−3654
−3656
−3658
0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24 0.26 0.28
Spacing [Å]
difference between 0.12 Å and 0.10 Å spacing is about 0.18 eV. It is very
small with respect the energy: only 0.005% of the energy. And in addition,
the time of the calculation is important. So, we choose 0.12 Å spacing for
calculation of pentacene.
We do the same calculation for picene. We can see the result in Fig. (3.2).
We can see that the suitable spacing for picene is the same as for pentacene.
−3640
−3641
Energy [eV]
−3642
−3643
−3644
−3645
0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Spacing [Å]
We can compare the results for picene and pentacene in Fig. (3.3). The
spacing just depends on a type of atoms that compose the molecules. In fact
it depends on the pseudopotential. So for picene and pentacene we have the
same suitable spacing that is equal to 0.12 Å.
3.1. GROUND STATE PROPERTIES 25
−3646
−3648
−3650
−3652
−3654
−3656
−3658
0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24 0.26 0.28
Spacing [Å]
−3648
Energy [eV]
−3650
−3652
−3654
−3656
−3658
2 2.5 3 3.5 4 4.5 5
Radius [Å]
bigger than 4 Å, we have very small error. For example the difference
between 4 Å and 5 Å radius is about 0.01 eV. It is very small with respect
the energy: only 3 × 10−4 % of the energy. And in addition, the time of the
calculation is important. So, we choose 4 Å radius for the calculation on
pentacene.
26 CHAPTER 3. RESULTS
We do the same calculation for picene. We can see the result in Fig. (3.5).
We can see that the suitable radius for picene is the same as for pentacene.
−3628
−3630
Energy [eV]
−3632
−3634
−3636
−3638
−3640
2 2.5 3 3.5 4 4.5 5
Radius [Å]
We can compare the results for picene and pentacene in Fig. (3.6). The
−3630
−3635
Energy [eV]
−3640
−3645
−3650
−3655
−3660
2 2.5 3 3.5 4 4.5 5
Radius [Å]
Figure 3.6: Convergence with radius of box for picene and pentacene
radius of the simulation box just depends on the type of atoms that compose
the molecules. In fact it depends on the pseudopotential. So for picene and
pentacene we have the same suitable radius that is equal to 4 Å.
I found that the minimisation was very slow. For this reason, we decided
to use the structures that appear in the (American) National Institute of
Standard and Technology (NIST) website.
From (http://webbook.nist.gov/cgi/cbook.cgi?Str3File=C135488) for pen-
tacene and for picene from (http://webbook.nist.gov/cgi/cbook.cgi?ID=213-
46-7).
In Fig. (3.7(a)) and Fig. (3.7(b)) we show the optimisation structures of
pentacene and picene.
As can be inferred from Fig. (3.7(a)), pentacene is symmetric with respect
to the xy, xz, and yz planes. While picene (see Fig. 3.7(b)) is symmetric
only with respect to the yz and xy plane. Thus picene has fewer symmetries
than pentacene.
Now, we compare C–C and C–H bond lengths in pentacene, picene, and
benzene.
All C–H bond lengths in picene, pentacene, and benzene are equal to 1.10
Å. All C–C bond lengths in benzene are equal to 1.40 Å, but in picene and
pentacene we can see different bond lengths. For instance, we compare the
C–C bond lengths in the middle rings for pentacene and picene. One could
expect that the bond lengths of C–C atoms 6–7 in picene is equal to 7–16
in pentacene, we can see the bond length of 6–7 in picene is equal to 1.36 Å
but 7–16 in pentacene is equal to 1.43 Å. Another example is bond length
of C atoms 5–6 in picene and 7–6 in pentacene. We can see the bond length
of 5–6 in picene is equal to 1.42 Å and 7–6 in pentacene is equal to 1.40 Å.
So, we can see that we have different bond lengths in pentacene and picene.
In Table (3.2) we write the energy width of the Kohn-Sham levels that
belong to each group of states. While the σ states have the different energy
width (13.58 eV in pentacene and 13.92 eV in picene), the σ ∗ states present
smaller and different energy width (4.25 eV in pentacene and 4.42 eV in
picene). The energy width of π states is larger for pentacene than for picene.
It is 3.71 eV for π states and 3.95 eV for π ∗ states in pentacene, while it is
2.76 eV for π states and 4.42 eV for π ∗ states in picene.
In Table (3.3) we compare the energy gap between the highest occu-
pied molecular orbital (HOMO) and lowest unoccupied molecular orbital
(LUMO) in picene and pentacene. In our calculation the HOMO–LUMO
energy gap for pentacene is 1.13 eV. We compare it with the data that we
find in literature.
3.1. GROUND STATE PROPERTIES 29
Pentacene Picene
σ band(eV) 13.58 13.92
π band(eV) 3.71 2.76
π ∗ band(eV) 3.95 2.98
σ ∗ band(eV) 4.25 4.42
Pentacene Picene
HOMO–LUMO energy gap (eV) 1.13 3.09
The agreement with other calculations is very good: in Ref. [30] the authors
find 1.1 eV and in Ref. [31] they find 1.12 eV. From this comparison, we
understand that our calculation is correct.
But when we compare this value with experimental data, we see that this
value is about 80% smaller than the experimental value, 5.22 eV [30]. The
underestimation of the energy gap is due to two reasons. First, because the
Kohn-Sham orbitals have no physical meaning. Second, because when we
model Exc using a local density approximation where the system is locally
approximated by the homogeneous electron gas of density n = n(r).
The same observation holds for picene. For picene in our calculation the
HOMO–LUMO energy gap is 3.09 eV. In Ref. [31] the authors report a value
of 2.96 eV. So, again, we conclude that our calculation is correct.
Figure 3.8: Kohn-Sham orbital for picene (a) Kohn-Sham orbital of the
HOMO (b) Kohn-Sham orbital of the LUMO (c) Kohn-Sham orbital of the
last σ state (d) Kohn-Sham orbital of the first σ ∗ state
32 CHAPTER 3. RESULTS
Figure 3.9: Kohn-Sham orbital for pentacene (a) Kohn-Sham orbital of the
HOMO (b) Kohn-Sham orbital of the LUMO (c) Kohn-Sham orbital of the
last σ state (d) Kohn-Sham orbital of the first σ ∗ state
3.2. ABSORPTION SPECTRA FROM TIME PROPAGATION 33
5
Absorption [Å ]
2
−1
0 5 10 15 20
Energy [eV]
Figure 3.10: The absorption spectrum for pentacene in one direction for
TDDeltaStrength = 0.005 (1/Å) and TDDeltaStrength = 0.01 (1/Å)
We should choose the suitable length of each time step. Longer time steps
are desirable to assure a shorten computational propagation time, but they
limits to the maximum frequency we will able to observe. So, we should find
the largest value that it is possible without the evolution becomes unstable.
For this purpose, we should run the calculation just with the time-evolution
of the system, departing from the ground-state, under the influence of no
external perturbation. As a consequence if the time step is correct, the
electronic system evolve in such a way that the total energy does not change.
We do this calculation with the total time propagation equal to 0.1 (}/eV )
and time step equal to 0.002 (}/eV ), 0.0015 (}/eV ), and 0.001 (}/eV ). We
found that the total energy of the system is conserved just for the time step
equal to 0.001 (}/eV ). So, we chose the time step equal to 0.001 (}/eV ) for
our calculation.
For checking that this time step is correct, we did the calculation also
in presence of an external perturbation. We did the calculation with ex-
ternal perturbation in one direction with total time propagation equal to
15 (}/eV ), and compare the result for time step equal to 0.001 (}/eV ) with
that for time step equal to 0.0005 (}/eV ).
In Fig. (3.11) we can see the absorption spectrum for two time steps for
pentacene and in Fig. (3.12) for picene. In these two figures, we plot the
strength function Eq. (2.39).
We can see that we have the same results for both time steps. From these
12000
StrengthFunction [1/eV]
10000
8000
6000
4000
2000
−2000
0 5 10 15 20
Energy [eV]
5000
4000
3000
2000
1000
−1000
0 5 10 15 20
Energy [eV]
Increasing the total propagation time (T) reduces the width of the peaks.
We do the calculation for time step equal to 0.001 (}/eV ) and T=10 (}/eV ),
T=20 (}/eV ), T=25 (}/eV ), T=30 (}/eV ), T=35 (}/eV ), and T=40 (}/eV ).
In Fig. (3.13) for pentacene and in Fig. (3.14) for picene, we plot the average
absorption coefficient T r σ/3 for these different total propagation time. As
we can see, when we increase the total propagation time, the width of the
peaks reduces. Moreover the increase of total propagation time causes the
appearance of new structures in the spectra indicating that we are not in
convergence up to about T=40 (}/eV ). We checked that above this value
the spectra does not change any more.
Thus, we find the suitable total propagation time equal to 40 (}/eV ).
First we start with the analysis of the pentacene spectrum. In Fig. (3.15(a))
we show the absorption spectrum for pentacene obtained plotting the av-
erage absorption coefficient while in Fig. (3.15(b)) we show the absorption
spectrum for three different direction of polarizations obtained from the
36 CHAPTER 3. RESULTS
Absorption [Å ]
12
2
10
8
6
4
2
0
−2
0 2 4 6 8 10 12
Energy [eV]
Figure 3.13: The absorption spectrum for (T=10 (}/eV ), T=20 (}/eV ),
T=25 (}/eV ), T=30 (}/eV ), T=35 (}/eV ), and T=40 (}/eV )) for pen-
tacene
strength function.
According with the distribution of the Kohn-Sham energy levels shown in
Table (3.2) and Table (3.3) we can distinguish two main energy windows
in the absorption spectra: a low energy region between 0 eV and 4.84 eV
where only π−π ∗ transitions are allowed by energy conservation and a higher
energy one (above 4.84 eV) involving both π − π ∗ and π − σ ∗ (σ − π ∗ ) transi-
tions. As can be inferred from Fig. (3.15(b)) the lower energy region of the
spectrum is dominated by electronic transitions induced by in-plane dipole
moment, in particular in the x direction. While the out-of-plane component
(z component) of the dipole moment gives a remarkably contribution only at
energy higher than 10 eV where only π −σ ∗ (σ −π ∗ ) transitions are involved.
This behaviour can be explained in terms of the symmetry properties of the
electronic wave functions discussed in the previous section. In fact as we
have seen previously π (π ∗ ) and σ (σ ∗ ) states are respectively antisymmetric
and symmetric respect to the xy plane. This means that the only allowed
electronic transitions in the dipole approximation are π − π ∗ and σ − σ ∗ for
polarization direction belonging to the plane of the molecule and π − σ ∗ and
σ − π ∗ for the direction perpendicular to the plane. Interestingly, we note
that, although the main contribution to the lower energy region of the spec-
trum arises from the x component of the polarization (see for example the
feature at 3.98 eV), the onset of the absorption (1.58 eV) involving mainly
HOMO-LUMO transition is related only to the y component. In fact being
HOMO and LUMO orbitals both symmetric respect to the yz plane, the
HOMO-LUMO transition is forbidden for dipole moment along the x axis.
On the other hand it is allowed for dipole along the y axis being HOMO and
LUMO orbitals respectively antisymmetric and symmetric respect to the xz
plane.
3.2. ABSORPTION SPECTRA FROM TIME PROPAGATION 37
5
4
3
2
1
0
−1
0 2 4 6 8 10 12
Energy [eV]
Figure 3.14: The absorption spectrum for (T=10 (}/eV ), T=20 (}/eV ),
T=25 (}/eV ), T=30 (}/eV ), T=35 (}/eV ), and T=40 (}/eV )) for picene
20
(a)
15
Absorption [Å ]
2
10
0
0 5 10
Energy [eV]
40000
StrengthFunction [1/eV]
x direction
30000 (b) y direction
z direction
20000
10000
0
0 5 10
Energy [eV]
Figure 3.15: The absorption spectrum for pentacene (a)The average absorp-
tion (b)The strength function for three direction (x,y,z)
8 (a)
Absorption [Å ]
2
6
0
0 5 10
Energy [eV]
20000
StrengthFunction [1/eV]
(b) x direction
15000 y direction
z direction
10000
5000
0
0 5 10
Energy [eV]
Figure 3.16: The absorption spectrum for picene (a)The average absorption
(b)The strength function for three direction (x,y,z)
StrengthFunction [1/eV]
4000
3000
2000
1000
0
0 2 4 6 8 10 12
Energy [eV]
Figure 3.17: The absorption spectrum for Casida equation with the number
of unoccupied states are equal to 51, 100, 150, and 200 for pentacene
3000
2500
2000
1500
1000
500
0
0 2 4 6 8 10 12
Energy [eV]
Figure 3.18: The absorption spectrum for Casida equation with the number
of unoccupied states are equal to 51, 100, and 150 for picene
difference of about 3.47 eV. Thus the Hartree term causes a red shift of the
onset of about 0.45 eV respect to the value set by the HOMO-LUMO gap.
The dominant peak at 3.98 eV is related to the mixing of transitions with
energy from 2.899 eV to 4.98 eV revealing a strong effects of the Hartree
term.
Now we focus on picene.
The onset of the absorption is at about 3.4 eV. We find from the solution of
Casida equation that this excitation is related to the mixing of transitions
with energy from 3.09 eV to 4.184 eV. Thus the Hartree term causes a red
shift of the onset of about 0.31 eV respect to the value set by the HOMO-
LUMO gap (3.09 eV).
The dominant peak at 4.39 eV is related to the mixing of transitions with
3.3. ABSORPTION SPECTRA FROM CASIDA EQUATION 41
6000
StrengthFunction [1/eV]
5000
4000
(a)
3000
2000
1000
0
0 5 10
Energy [eV]
20
(b)
15
Absorption [Å ]
2
10
0
0 5 10
Energy [eV]
Figure 3.19: The absorption spectrum for pentacene (a) absorption spec-
trum from Casida (b) the absorption spectrum from time propagation
energy from 3.09 eV to 6.06 eV. Even in this case the mixing of transitions
due to the Hartree term is very strong.
42 CHAPTER 3. RESULTS
5000
StrengthFunction [1/eV]
(a)
4000
3000
2000
1000
0
0 5 10
Energy [eV]
8
(b)
Absorption [Å ]
2
0
0 5 10
Energy [eV]
Figure 3.20: The absorption spectrum for picene (a) absorption spectrum
from Casida (b) the absorption spectrum from time propagation
4000
StrengthFunction [1/eV]
3500
3000
2500
2000
1500
1000
500
0
0 2 4 6 8 10 12
Energy [eV]
Figure 3.21: The absorption spectrum for Casida equation with fxc and
without fxc of pentacene
3.3. ABSORPTION SPECTRA FROM CASIDA EQUATION 43
2500
2000
1500
1000
500
0
0 2 4 6 8 10 12
Energy [eV]
Figure 3.22: The absorption spectrum for Casida equation with fxc and
without fxc of picene
Chapter 4
Conclusion
45
46 CHAPTER 4. CONCLUSION
both HOMO and LUMO orbitals are symmetric respect to the yz plane
while respect to the xz plane they are anti-symmetric and symmetric re-
spectively. On the other hand, for picene (see Fig. (3.8(a)) and (3.8(b)))
the HOMO and LUMO orbitals are symmetric and anti-symmetric respect
to the yz plane respectively but they do not have any symmetry respect to
the xz plane. Moreover being the HOMO (LUMO) orbitals as well as all
the π (π ∗ ) states linear combinations of pz atomic orbitals they are anti-
symmetric respect to the plane of the molecule (xy plane) in both picene
and pentacene. In Fig. (3.8(c)) and (3.9(c)) with Fig. (3.9(d)) and (3.8(d))
we see σ and σ ∗ states that they are symmetric respect the molecular plane.
Then we find the absorption spectrum with two ways: first from time
propagation and second from Casida equation. In the spectra that we ob-
tained from time propagation for pentacene (Fig. (3.15)) and for picene
(Fig. (3.16)) we have two main energy windows: a low energy region be-
tween 0 eV and 4.84 eV for pentacene and for picene between 0 eV and 5.85
eV where only π − π ∗ transitions are allowed by energy conservation and a
higher energy one (above 4.84 eV for pentacene and for picene above 5.85
eV) involving both π − π ∗ and π − σ ∗ (σ − π ∗ ) transitions. The lower energy
region of the spectrum is dominated by electronic transitions induced by
in-plane dipole moment. The main contribution to the lower energy region
of the spectrum arises from the x component of the polarization for picene
and pentacene. The onset of the absorption for pentacene involving mainly
HOMO-LUMO transition is related only to the y component, but for picene
the onset of the absorption presents contributions from both the x compo-
nent and the y component of the dipole moment.
We see the spectra from Casida with that obtained from time propagation
are in good agreement. We see that the effect of fxc is negligible in the
first region of the spectra and causes only a small red shift of the spectra
of about 0.1 eV in both picene (Fig. (3.22)) and pentacene (Fig.(3.21)) at
high energy. This means that in these systems the mixing of electron-hole
transitions is mainly related to the Hartree term.
Bibliography
47
48 BIBLIOGRAPHY
[22] E. Runge and E. K. U. Gross, Phy. Rev. Lett. 52, 997 (1984).
[29] A. Castro et al., Phys. Stat. Sol. (b) 243, 2465 (2006).
[31] T. Kosugi et al., Journal of the Physical Society of Japan 78, 113704
(2009).
List of Figures
49
50 LIST OF FIGURES
51