Sei sulla pagina 1di 5

Electronic transport in organometallic perovskite CH3NH3PbI3: The role of organic

cation orientations
G. R. Berdiyorov, F. El-Mellouhi, M. E. Madjet, F. H. Alharbi, and S. N. Rashkeev
Citation: Applied Physics Letters 108, 053901 (2016); doi: 10.1063/1.4941296
View online: http://dx.doi.org/10.1063/1.4941296
View Table of Contents: http://scitation.aip.org/content/aip/journal/apl/108/5?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Efficient thermal conductance in organometallic perovskite CH3NH3PbI3 films
Appl. Phys. Lett. 108, 081902 (2016); 10.1063/1.4942779
Charge transport in bulk CH3NH3PbI3 perovskite
J. Appl. Phys. 119, 074101 (2016); 10.1063/1.4941532
A simple rule for determining the band offset at CH3NH3PbI3/organic semiconductor heterojunctions
Appl. Phys. Lett. 108, 021602 (2016); 10.1063/1.4939744
Atomistic origins of CH3NH3PbI3 degradation to PbI2 in vacuum
Appl. Phys. Lett. 106, 131904 (2015); 10.1063/1.4916821
Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber
Appl. Phys. Lett. 104, 063903 (2014); 10.1063/1.4864778

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 103.21.127.78 On: Sun, 09 Oct 2016
13:43:34

APPLIED PHYSICS LETTERS 108, 053901 (2016)

Electronic transport in organometallic perovskite CH3NH3PbI3:


The role of organic cation orientations
G. R. Berdiyorov,1,a) F. El-Mellouhi,1 M. E. Madjet,1 F. H. Alharbi,1,2 and S. N. Rashkeev1
1

Qatar Environment and Energy Research Institute, Hamad Bin Khalifa University, Qatar Foundation, Doha,
Qatar
2
College of Science and Engineering, Hamad Bin Khalifa University, Doha, Qatar

(Received 19 October 2015; accepted 22 January 2016; published online 4 February 2016)
Density functional theory in combination with the nonequilibrium Greens function formalism is
used to study the electronic transport properties of methylammonium lead-iodide perovskite
CH3NH3PbI3. Electronic transport in homogeneous ferroelectric and antiferroelectric phases, both
of which do not contain any charged domain walls, is quite similar. The presence of charged domain wall drastically (by about an order of magnitude) enhances the electronic transport in the lateral direction. The increase of the transmission originates from the smaller variation of the
electrostatic potential profile along the charged domain walls. This fact may provide a tool for tuning transport properties of such hybrid materials by manipulating molecular cations having dipole
C 2016 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4941296]
moment. V

Organometallic lead-halide perovskites, such as methylammonium (MA) lead iodide (CH3NH3PbI3 or MAPbI3),
have attracted new interest in the past few years as promising
light harvesting materials with the solar conversion efficiency already exceeding 20%1 (see Refs. 27 for review).
Such an excellent performance of perovskite solar cells is
mainly related to: strong light absorption, proper band structure and band alignment,8 and relatively long carrier diffusion length.9,10 The properties of MAPbI3 can further be
refined by chemical engineering, e.g., through mixing halogen atoms.1117 Moreover, because of the tolerance of these
materials to topological defects,18 perovskite based solar
cells can be fabricated using low-temperature solution methods, which result in low manufacturing cost.
The organic cations play an important role in determining the structural, electronic, and optical properties of organometallic perovskites.1922 This is because the dynamic
reorientation of the organic molecules disturbs the metalhalide lattice23,24 affecting the properties of the material. For
example, orientational rotation of the cations results in
dynamic direct-to-indirect transition in the band gap, which
suppresses considerably the charge carrier recombination.22
The charged nature of the organic molecules can also contribute to the superior solar conversion efficiency of leadhalide perovskites.2429 For example, ferroelectric domain
walls can be formed due to the permanent dipole moment of
the MA molecules,26,29 which can assist the separation of
electrons and holes and reduce their recombination due to
charge segregation.21,25,30 The conductivity along the
charged domains can also be significantly increased due to
the charge carrier accumulation.10 Despite recent extensive
research, the role of organic cations in defining the excellent
photovoltaic performance of these materials is not fully
understood.

a)

Electronic mail: gberdiyorov@qf.org.qa

0003-6951/2016/108(5)/053901/4/$30.00

In this work, we used first-principles density-functional


theory (DFT) in combination with the nonequilibrium
Greens function formalism to study the electronic transport
properties of MAPbI3 in different ferroelectric states. We
consider the cubic crystalline phase of MAPbI3 in order to
reduce the structural changes induced by the organic base.31
We found that the electronic transport in the system can be
increased by an order of magnitude along the charged domain walls. Such an enhancement of electronic transport
originates from smaller variations of the electrostatic potential profile along the domain walls, whereas the effect of internal octahedral reorganization is negligibly small in this
crystal phase of the material.
As a typical example, we consider cubic phase of
MAPbI3 with MA molecules oriented along the 100 direction
relative to the origin of the cubic lattice. Hereafter, we call
this ferroelectric arrangement as sample 1 [see Fig. 1(a)]. By
altering the orientation of MA molecules in neighboring cells,
we create an antiferroelectric state [sample 2, Fig. 1(b)]. Such
ferroelectric and antiferroelectric states have been observed in
recent low-temperature scanning tunneling microscopy studies of methylammonium lead halide perovskites.32 In the last
system [sample 3, Fig. 1(c)], we have created charged domain
walls by reorienting the MA molecules. All the samples were
geometry optimized using DFT within the generalized gradient approximation (GGA) in the Perdew-Burke-Ernzerhof
(PBE) form to represent the exchange-correlation energy.33
The Brillouin zone integration was performed using 6  6  6
k-point sampling.34 Grimmes DFT-D3 empirical dispersion
correction35 to the PBE was used to account for van der
Waals interactions, which is known to be very important for
an accurate description of the structural properties of organicinorganic hybrid materials.19,36 Calculations are performed
using the Vienna ab initio simulation package (VASP).37,38
,
The calculated lattice constant for sample 1 was 6.315 A
39

which is close to the experimental value of 6.33 A. Since


the simulation box is not relaxed during the optimization

108, 053901-1

C 2016 AIP Publishing LLC


V

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 103.21.127.78 On: Sun, 09 Oct 2016
13:43:34

053901-2

Berdiyorov et al.

Appl. Phys. Lett. 108, 053901 (2016)

FIG. 2. (a) Zero bias transmission spectra and (b) device density of states as
a function of electron energy (the Fermi energy is taken at E 0) for all considered samples. Thin curves show the results obtained for the samples optimized by fixing the positions of lead and iodine atoms.

FIG. 1. Device geometries: cubic lattice structure of organometallic perovskite CH3NH3PbI3 for different orientations of the organic molecules: (a)
sample 1, (b) sample 2 and (c) sample 3. Periodic boundary conditions are
applied along the x- and y-directions and electronic transport occurs along
the z-direction through the metallic electrodes. Arrows indicate the direction
of the dipole moment of the organic cations.

(only the positions of atoms are changed), the lattice parameters of the other two samples remain the same.
Lattice constants and atomic coordinates from the
VASP calculations were used to calculate the electronic
transport properties of the samples using the first-principles
computational package Atomistix toolkit (ATK).40 The electronic properties are described self-consistently within the
DFT/GGA, and the electrostatic potentials were determined
on a real-space grid with a cutoff energy of 150 Ry. Doublezeta-double-polarized basis sets of local numerical orbitals
were employed for all atoms. To calculate the electronic
transport through the system, we have constructed device
geometries (see Fig. 1), each of which consists of left and
right regions that connect a central region (i.e., two-probe
configuration). At the left and right parts of the sample we
put metallic electrodes, with in plane periodic boundary conditions perpendicular to the sample-metal contact. Quantum
transport is calculated using the nonequilibrium Greens
function with the 6  6  100 k-point sampling.
As a main result, we present in Fig. 2(a) the zero bias
transmission spectra, T(E), as a function of the electronic
energy for all the considered samples. We show mostly the
region near the Fermi level, which is the energy origin in the
T(E) plots, since for small bias voltages this is the relevant
energy range that contributes to electronic transport. In the
case of ferroelectric sample 1, zero transmission is obtained in
the energy range [0.78, 0.78] eV due to the energy band gap
[solid black curve in Fig. 2(a)]. The estimated band gap of

1.56 eV is close to the experimental value 1.55 eV,41 and previous DFT-GGA result 1.57 eV.19 Below the Fermi level, the
transmission spectrum is characterized by well separated
peaks with different amplitudes. The transmission peaks
become less pronounced above the Fermi level. These features are also reflected in the device density of states (DDOS)
of the system as sharp peaks below the Fermi energy and
plateau-like behavior at larger energies [see the solid-black
curve in Fig. 2(b)], indicating that there is a correspondence
between the energy levels of the perovskite material and the
transmission spectrum. Interestingly, the transmission spectrum does not change much when the system is turned into
the antiferroelectric state [dashed red curve in Fig. 2(a)]: T(E)
increases slightly near the band edges due to small reduction
of the band gap (0.1 eV). These two samples also show similar DDOS [Fig. 2(b)].
The transmission spectrum changes considerably when
the MA molecules form a charged domain wall [dotted blue
curve in Fig. 2(a)]. There are extra features in the spectrum
which are different from those obtained for the other two
samples. For example, the region where transmission is zero
becomes narrower due to the 0.5 eV reduction of the electronic band gap; the value of this reduction is in agreement
with recent DFT calculations.26 Because of the smaller band
gap, sample 3 has larger transmission near the Fermi level,
which results in larger current at small bias voltages. Finally,
the transmission spectrum of sample 3 shows energy dependent asymmetry, with larger transmission for electron energies
above the Fermi level. The values of the transmission inside
the valence band in sample 3 are lower than those in samples
1 and 2, despite of larger DDOS in sample 3 in this energy
range [see dotted blue curve in Fig. 2(b)]. Thus, the presence
of charged domain wall significantly changes the transport
properties of MAPbI3 samples. Notice that the electronic
transmission across the domain walls is significantly lower
(see the zero bias transmission spectra for samples 5 and 6 in
supplementary Fig. S242).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 103.21.127.78 On: Sun, 09 Oct 2016
13:43:34

053901-3

Berdiyorov et al.

Appl. Phys. Lett. 108, 053901 (2016)

FIG. 3. Current-voltage characteristics


of MAPbI3 for different orientations of
the MA molecules.

It is known that the Pb-I chemistry plays an important


role in determining the electronic properties of the leadhalide perovskites. Certain orientations of the organic cations
can affect the corresponding Pb-I inorganic frame, altering
the transport properties of the material.22,31,43 To exclude the
contribution of the structural changes of the Pb-I lattice to
the transmission spectra, we have optimized samples 2 and 3
by keeping the positions of the lead and iodine atoms
unchanged. The transmission spectra of these new structures
are shown in Fig. 2(a) by thin curves. In both cases, the
changes in the transmission spectrum due to the Pb-I relaxation are negligibly small, especially near the band edges. A
small increase in the band gap of the material is observed
due to the relaxation of the inorganic frame. Thus, the
changes in the electronic transmission due to the reorientation of the MA molecules originate mainly from charged organic cations with dipole moment.
To study the response of the considered samples to finite
voltage biasing, we have calculated the current-voltage (I-V)
characteristics of these materials using the LandauerB
uttiker formula

2e lR

(1)
T E; V f E  lL  f E  lR dE;
I V
h lL
where f E; EF is the Fermi-Dirac distribution function and
lL/lR is the chemical potential of the left/right electrode.
Figure 3 shows the I-V curves of the samples for the bias
voltage below 2 V. It is clear that in samples without charged
domain walls, ferroelectric ordering has no substantial
impact on the transport properties of the material, and the
current generated in the antiferroelectric sample is just a bit
smaller than that in the ferroelectric one (compare filled
black and open red circles in Fig. 3). The situation changes
considerably when the charged domains are present in the
system (filled blue squares in Fig. 3). Already, starting from
the low bias voltages, the current in sample 3 shows larger
electronic transport as compared with the other two samples.
In fact, the current in sample 3 can be an order of magnitude
larger for these values of the bias voltage. The difference in
electronic transport is also obtained for larger values of the
applied voltage. Open-green squares in Fig. 3 show the
results obtained for the sample with larger unit cell (sample
4, see the inset of Fig. 3). This system shows higher current
density as compared with that of sample 3. Notice that

sample 4 has the smallest formation energy among all the


considered samples (see Fig. S1 and Table I in the supplementary material42). These findings indicate an important
role of charged domain walls in transport properties of organometallic perovskites.
Charge density localization is one of the important factors
affecting the transport of charge carriers in perovskite materials.30 To reveal the features of electron localization in our systems, we have calculated the electron localization functions
(ELFs), which are shown in Fig. 4 for all the considered samples. The results are shown along the Pb-I planes (right column) and along the plane of the MA molecules (left column).
Strong localization of electrons is obtained around the lead
and iodine atoms with an overlap of orbitals along the Pb-I
plane. Localization of electrons near the MA molecules is
also obtained. However, the charge localization is not affected
much by the domain walls: similar ELF is obtained for all
three samples, especially along the Pb-I plane. So, the
improvement in the electronic transport should be attributed
to other factors, such as the electrostatic potential profile.
Figure 5 shows the averaged electrostatic difference potential

FIG. 4. Contour plots of the electron localization function for sample 1 (a)
and (b), sample 2 (c) and (d), and sample 3 (e) and (f). The results are shown
along the Pb-I plane (right column) and along the plane of the MA molecules (left column).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 103.21.127.78 On: Sun, 09 Oct 2016
13:43:34

053901-4

Berdiyorov et al.

Appl. Phys. Lett. 108, 053901 (2016)


4

P. Gao, M. Gratzel, and M. K. Nazeeruddin, Energy Environ. Sci. 7, 2448


(2014).
H. Zhou, Q. Chen, G. Li, S. Luo, T.-b. Song, H.-S. Duan, Z. Hong, J. You,
Y. Liu, and Y. Yang, Sience 345, 542 (2014).
6
Y.-C. Hsiao, T. Wu, M. Li, Q. Liu, W. Qin, and B. Hu, J. Mater. Chem. A
3, 15372 (2015).
7
J. Berry, T. Buonassisi, D. A. Egger, G. Hodes, L. Kronik, Y.-L. Loo, I.
Lubomirsky, S. R. Marder, Y. Mastai, J. S. Miller, D. B. Mitzi, Y. Paz, A.
M. Rappe, I. Riess, B. Rybtchinski, O. Stafsudd, V. Stevanovic, M. F.
Toney, D. Zitoun, A. Kahn, D. Ginley, and D. Cahen, Adv. Mater. 27,
5102 (2015).
8
P. Umari, E. Mosconi, and F. De Angelis, Sci. Rep. 4, 4467 (2014).
9
S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. Alcocer, T.
Leijtens, L. M. Herz, A. Petrozza, and H. J. Snaith, Science 342, 341 (2013).
10
S. N. Rashkeev, F. El-Mellouhi, S. Kais, and F. H. Alharbi, Sci. Rep. 5,
11467 (2015).
11
J. H. Noh, S. H. Im, J. Heo, T. N. Mandal, and S. I. Seok, Nano Lett. 13,
1764 (2013).
12
A. Amat, E. Mosconi, E. Ronca, C. Quarti, P. Umari, M. K. Nazeeruddin,
M. Graetzel, and F. De Angelis, Nano Lett. 14, 3608 (2014).
13
G. Xing, N. Mathews, S. S. Lim, N. Yantara, X. Liu, D. Sabba, M.
Gratzel, S. Mhaisalkar, and T. C. Sum, Nat. Mater. 13, 476 (2014).
14
N. J. Jeon, J. H. Noh, W. S. Yang, Y. C. Kim, S. Ryu, J. Seo, and S. Seok,
Nature (London) 517, 476 (2015).
15
W. Nie, H. Tsai, R. Asadpour, J.-C. Blancon, A. J. Neukirch, G. Gupta, J.
J. Crochet, M. Chhowalla, S. Tretiak, M. A. Alam, H.-L. Wang, and A. D.
Mohite, Science 347, 522 (2015).
16
D. Shi, V. Adinolfi, R. Comin, M. Yuan, E. Alarousu, A. Buin, Y. Chen,
S. Hoogland, A. Rothenberger, K. Katsiev, Y. Losovyj, X. Zhang, P. A.
Dowben, O. F. Mohammed, E. H. Sargent, and O. M. Bakr, Science 347,
519 (2015).
17
D. W. deQuilettes, S. M. Vorpahl, S. D. Stranks, H. Nagaoka, G. E. Eperon,
M. E. Ziffer, H. J. Snaith, and D. S. Ginger, Science 348, 683 (2015).
18
W. J. Yin, T. Shi, and Y. Yan, Appl. Phys. Lett. 104, 063903 (2014).
19
E. Mosconi, A. Amat, Md. K. Nazeeruddin, M. Gratzel, and F. De
Angelis, J. Phys. Chem. C 117, 13902 (2013).
20
F. Brivio, A. B. Walker, and A. Walsh, APL Mater. 1, 042111 (2013).
21
E. Mosconi, C. Quarti, T. Ivanovska, G. Ruani, and F. De Angelis, Phys.
Chem. Chem. Phys. 16, 16137 (2014).
22
C. Motta, F. El-Mellouhi, S. Kais, N. Tabet, F. Alharbi, and S. Sanvito,
Nat. Commun. 6, 7026 (2015).
23
T. Baikie, N. S. Barrow, Y. Fang, P. J. Keenan, P. R. Slater, R. O. Piltz,
M. Gutmann, S. G. Mhaisalkar, and T. J. White, J. Mater. Chem. A 3,
9298 (2015).
24
A. M. A. Leguy, J. M. Frost, A. P. McMahon, V. G. Sakai, W.
Kochelmann, C. Law, X. Li, F. Foglia, A. Walsh, B. C. ORegan, J.
Nelson, J. T. Cabral, and P. R. F. Barnes, Nat. Commun. 6, 7124 (2015).
25
J. M. Frost, K. T. Butler, F. Brivio, C. H. Hendon, and M. van
Schilfgaarde, Nano Lett. 14, 2584 (2014).
26
S. Liu, F. Zheng, N. Z. Koocher, H. Takenaka, F. Wang, and A. M. Rappe,
J. Phys. Chem. Lett. 6, 693 (2015).
27
Z. Fan, J. Xiao, K. Sun, L. Chen, Y. Hu, J. Ouyang, K. P. Ong, K. Zeng,
and J. Wang, J. Phys. Chem. Lett. 6, 1155 (2015).
28
Z. Fan, K. Sun, and J. Wang, J. Mater. Chem. A 3, 18809 (2015).
29
A. Walsh, J. Phys. Chem. C 119, 5755 (2015).
30
J. Ma and L.-W. Wang, Nano Lett. 15, 248 (2015).
31
F. Zheng, H. Takenaka, F. Wang, N. Z. Koocher, and A. M. Rappe,
J. Phys. Chem. Lett. 6, 31 (2015).
32
R. Ohmann, L. K. Ono, H.-S. Kim, H. Lin, M. V. Lee, Y. Li, N.-G. Park,
and Y. Qi, J. Am. Chem. Soc. 137, 16049 (2015).
33
J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
34
H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).
35
S. Grimme, J. Comput. Chem. 27, 1787 (2006).
36
D. A. Egger and L. Kronik, J. Phys. Chem. Lett. 5, 2728 (2014).
37
G. Kresse and J. Furthmuller, Comput. Mater. Sci. 6, 15 (1996).
38
G. Kresse and D. Joubert, Phys. Rev. B 59, 1758 (1999).
39
A. Poglitsch and D. Weber, J. Chem. Phys. 87, 6373 (1987).
40
See http://www.quantumwise.com for the QuantumWise company,
Copenhagen, Denmark.
41
A. Kojima, K. Teshima, Y. Shirai, and T. Miyasaka, Am. Chem. Soc. 131,
6050 (2009).
42
See supplementary material at http://dx.doi.org/10.1063/1.4941296 for
optimized structures, total energy values, transmission spectra and electrostatic potential variations of all considered samples.
43
C. Motta, F. El Mellouhi, and S. Sanvito, Sci. Rep. 5, 12746 (2015).
5

FIG. 5. Electrostatic difference potential along the transport direction for


zero potential difference across the samples.

along the transport direction at zero voltage biasing for all the
considered samples. Periodic oscillations are obtained in all
systems with several peaks in each period. The largest oscillations of the electrostatic potential are obtained for the ferroelectric sample (solid black curve). The amplitude of the
oscillations becomes slightly smaller for the antiferroelectric
sample (dashed red curve). The scattering of the electrons
from these potential oscillations reduces the probability of
their transmission through the system. Significantly smaller
potential oscillations are observed for the sample with charged
domain walls (dotted blue curve). This increases the possibility of electron transfer across the sample and can explain the
enhanced electronic transport due to the charged domain
walls. The smallest amplitude of the potential variations is
obtained for sample 4 (see dashed dotted green curve), which
has resulted in largest electronic transport in this system (see
Fig. 3). The larger oscillations of the electrostatic potential are
also obtained when charged domain walls are oriented perpendicular to the direction of transport (see supplementary Fig.
S3 for samples 5 and 6), resulting in reduced electronic transmission (see supplementary Fig. S2).42
Summarizing, first-principles DFT calculations in combination with the nonequilibrium Greens function formalism
have been performed to study the effect of charged domain
walls on the electronic transport properties of lead-halide perovskite MAPbI3. We found that electronic transmission
increases considerably in the presence of charged domain
walls. For example, depending on the applied bias voltage, the
current in the direction parallel to the wall can be about an
order of magnitude higher as compared with the samples without domain walls. Such an enhancement of electronic transport
originates from the smaller variations of the electrostatic
potential profile along the domain walls. In the considered
cubic phase of the material, the structural changes due to the
reorientation of the organic molecules (e.g., Pb-I octahedral
tilts) have minor effect on the electronic transport properties of
the material. Our findings indicate the important role of ferroelectric domain walls in the charge carrier transport in organometallic perovskites. Computational resources were provided
by the research computing center at Texas A&M University in
Qatar.
1

M. A. Green, K. Emery, Y. Hishikawa, W. Warta, and E. D. Dunlop, Prog.


Photovoltaics Res. Appl. 23, 1 (2015).
2
M. A. Green, A. Ho-Baillie, and H. J. Snaith, Nat. Photonics 8, 506
(2014).
3
T. C. Sum and N. Mathews, Energy Environ. Sci. 7, 2518 (2014).

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. Download to IP: 103.21.127.78 On: Sun, 09 Oct 2016
13:43:34

Potrebbero piacerti anche