You are on page 1of 75

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/312304986

Perovskite Solar Absorbers and Ferroelectric Nanocomposites for Harvesting


Solar Energy

Research Proposal · April 2015


DOI: 10.13140/RG.2.2.31220.14724

CITATIONS READS
0 858

1 author:

Chaminda Hettiarachchi
University of South Florida
15 PUBLICATIONS   78 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Transparent Conducting Oxides View project

Quantum Dot Solar Cells View project

All content following this page was uploaded by Chaminda Hettiarachchi on 13 January 2017.

The user has requested enhancement of the downloaded file.


Perovskite Solar Absorbers and Ferroelectric Nanocomposites
For Harvesting Solar Energy

Chaminda Hettiarachchi

Dissertation Proposal for admission to candidacy


Laboratory for Advanced Materials Science and Technology
Department of Physics
University of South Florida
April 17, 2015

Co-Major Professor: Sarath Witanachchi, Ph. D.


Co-Major Professor: Pritish Mukherjee, Ph. D.
Hariharan Srikanth, Ph. D.
Denis Karaiskaj, Ph. D.
Manh-Huong Phan, Ph. D.
Abstract

Organometal halide perovskite absorbers such as methylammonium lead iodide chloride

(CH3NH3PbI3-xClx), have emerged as an exciting new material family for photovoltaics due to its

appealing features that include suitable direct bandgap with intense light absorbance, band gap

tunability, ultra-fast charge carrier generation, slow electron-hole recombination rates, long

electron and hole diffusion lengths, microsecond-long balanced carrier mobilities, and ambipolar

charge carrier mobilities. In a perovskite, absorption of photons results in the generation

of electron-hole pairs that evolve towards the formation of highly delocalized Wannier

t y p e excitons after thermalization. A fraction of which would dissociate spontaneously

back into free carriers. The excitons and free carriers co-exist and due to the crystalline

nature of the perovskite as well as low trap densities, the recombination within the

perovskites is limited. Ferroelectrics can be exploited as a driving force for charge separation

because spontaneous electric polarization promotes the desirable separation of photo-excited

carriers and allows a higher photovoltage and photocurrent, which may enable efficiencies

beyond the current maximum in perovskite solar cells. BaTiO3 nanoparticles is an excellent

ferroelectric material which has a high dielectric constant. A hybrid structure of CH3NH3PbI3-

xClx perovskites and ferroelectric BTO nanocomposite can be a new type of photovoltaic device.

The standard method of preparing CH3NH3PbI3-xClx perovskite precursors is a tedious process

involving multiple steps and, the chemicals being used (hydroiodic acid and methylamine) are

quite expensive. In this work we present a novel, single-step, simple, and cost-effective solution

approach to prepare CH3NH3PbI3-xClx thin films by the direct reaction of the commercially

available CH3NH3Cl (or MACl) and PbI2 .Thin film CH3NH3PbI3-xClx with enhanced

crystallinity have been produced by nebulizing precursors and a low-pressure spray deposition

2
technique. Optimum growth conditions have been identified. Deposition and characterization of

the hybrid ITO/CH3NH3PbI3-xClx:BTO/Al crystalline structure using the low-pressure spray

deposition technique is proposed. A balance between perovskite coverage and ferroelectric

nanoparticle concentration is a must for a more efficient device. For that, a systematic

investigation of the growth process that involves growth temperatures, BTO nanoparticle

concentrations, and film thicknesses have to be conducted. The final goal of the project is to

develop an understanding of the relation between the growth parameters and the efficiency of the

ferroelectric nanoparticles embedded perovskite solar cells.

3
TABLE OF CONTENTS

CHAPTER 1. DISSERTATION RESEARCH 1


1.1. Project Objectives 3
1.2.Research Plan 5

CHAPTER 2. BACKGROUND 6
2.1. Perovskite Solar Absorbers 7
2.2. Organometal Halide Perovskites 8
2.2.1. Methylammonium Lead Halide Perovskites 8
2.2.2. Crystal Structure 10
2.2.3. Temperature Dependent Structural Properties 12
2.2.4. Operational Principles of Perovskite Solar Cells 15
2.2.5. Wannier Type Excitons and Origin of a High Open-Circuit Voltage 16
2.3. Ferroelectric Nanocomposites 18
2.3.1. Ferroelectric Photovoltaic Effect 22
2.3.2. Bismuth Ferrite (BiFeO3 (BFO)) 23
2.3.3. Origin of the Large Photovoltage in FEPV Devices 24
2.3.4. PV properties of Pt/ BFO thin film/ FTO capacitor 26
2.3.5. Barium Titanate (BaTiO3 (BTO)) Nanoparticles 30
2.4. Low-Pressure Spray Deposition 33
2.5. Current Project Accomplishments 35
2.6. Project Challenges 35

CHAPTER 3. RESULTS 37
3.1. Synthesis of CH3NH3PbI3-xClx Perovskite 37
3.2. Deposition of CH3NH3PbI3-xClx Perovskite Films 38
3.3. Characterization of CH3NH3PbI3-xClx Perovskite Films 38
3.3.1. Structural Characterization 39
3.3.2. Optical Characterization 45

CHAPTER 4: MODEL TO GUIDE THE DEVICE PARAMETERS 47

APPENDIX 51
A.1 Schematic Diagram of the Vibrating-Mesh Nebulizer 51

REFERENCES 55
PUBLICATIONS & PRESENTATIONS 67

4
LIST OF FIGURES
Figure 1. Working principle of p-n junction solar cells. 1

Figure 2. Working principle of proposed ferroelectric nanocomposite/perovskite

photovoltaic device. 2

Figure 3. PV device architectures: (a) vertical and (b) lateral. 3

Figure 4. Crystalline systems of perovskites. 8


Figure 5. Crystal structure of Methylammonium lead halide perovskites. 11

Figure 6. Scheme ([006] view) of phase transitions of lead halide perovskite. 14

Figure 7. Open-circuit voltage versus optical bandgap 16

Figure 8. Typical hysteresis curve of a ferroelectric material. 19

Figure 9. Parallel and perpendicular configurations of photovoltage measurements

of BiFeO3 films. 25

Figure 10. Photovoltage vs light intensity of BFO films. 25

Figure 11. Dark and light illumination J–V characteristics for BFO films. 27

Figure 12. Illuminated J-V characteristics of BFO films in different

polarization states. 27

Figure 13. Ferroelectric polarization - electric field hysteresis loops measured at a

maximum applied electric field of 400 kV/cm and 10 kHz. 28

Figure 14. Energy band schematic diagrams of Pt/BFO film/FTO capacitor. 30

Figure 15. Curie temperature dependence on BTO particle size. 32

Figure 16. Experimental setup for the low-pressure spray deposition process. 34

Figure 17. Low-pressure spray deposition system. 34

Figure 18. X-ray diffraction patterns of fabricated CH3NH3PbI3-xClx perovskite films

at 6 different concentrations. 39

5
Figure 19. Top view SEM images of fabricated CH3NH3PbI3-xClx perovskite films

on glass substrates at different concentrations. 41

Figure 20. X-ray diffraction patterns CH3NH3PbI3-xClx perovskite films at

different deposition temperatures. 42

Figure 21. Top view of SEM images of fabricated CH3NH3PbI3-xClx perovskite films

on glass substrates at different deposition temperatures. 43

Figure 22. Linear absorption spectra of CH3NH3PbI3-xClx perovskite films for

6 different precursors. 44

Figure 23. Linear absorption spectra of CH3NH3PbI3-xClx perovskite films

at different deposition temperatures. 46

Figure. 24. Hypothetical cubic arrangement of BTO nanoparticles. 47

Figure. 25. Relationship between nanoparticle distance, size of a nanoparticle and

the distance from one nanoparticle to the center of the cube. 49

Figure. 26. Electric field versus distance from the center of a nanoparticle and

volume fraction of the absorber versus distance from

the center of a nanoparticle. 50

6
1. Dissertation Research
In traditional p–n junction solar cells (Fig. 1), the absorbed photons can pump the electrons from

the valence band of a light absorbing semiconductor material to its conduction band, with holes

left in the valence band. The photogenerated electrons and holes are quickly separated by the

built-in electric field inside the p–n junction and collected by the respective electrodes [135].

Theoretically, the magnitude of Voc in p–n junction solar cells is determined by the quasi-Fermi

energy difference of photogenerated electrons and holes which is limited by the bandgap of the

light absorbing semiconductors [135].

Fig. 1 working principle of p-n junction solar cells.

In this dissertation research, a novel device structure that consists of a perovskite solar absorber

and ferroelectric nanoparticles is investigated for solar energy harvesting. The newly discovered

perovskite phase of CH3NH3PbI3-xClx is currently under intense investigation by multiple groups

in the world for the unique properties it offer for solar energy harvesting. Devices of efficiency

up to 19.3 % have been reported. In this research, non p-n junction device is developed by

combining the efficient carrier generation of this material with possible electric fields generated

by ferroelectric nanoparticles. The rational for the structure is described below.

Spherical BaTiO3 ferroelectric nanoparticles possess a spontaneous permanent polarization P =

0.26 C m-2 [25]. Therefore, the local electric field due to this polarization around a nanoparticle

at a distance r can be written as

1

E BTO 
R3P
3 0 r 3
2 cosrˆ  sin ˆ 
where the ε0 is the free space permittivity. If the radius of a nanoparticle is R, the average electric

field magnitude at a distance of ξ = r - R = 5 nm (i.e., r = 30 nm) from the surface of the particle

is ~ 1010 Vm-1. Therefore there is a strong electric field close to the BTO nanoparticle surfaces

and when the light with a wavelength corresponding to the energy band gap is incident on

electrically poled BTO ferroelectric nanoparticles embedded CH3NH3PbI3-xClx perovskite,

photons are absorbed by the perovskite with the generation of charge carriers - electrons and

holes. The photogenerated electrons and holes are driven by the polarization-induced internal

electric field in opposite directions toward the cathode and anode, respectively, and thus

contribute to the photovoltaic output (Fig. 2).

Photons
Photo current

ITO top contact


electron

hole

Internal electric field


BTO
nanoparticles

Al bottom contact

Fig. 2 working principle of proposed ferroelectric nanoparticle/perovskite photovoltaic device.

A typical device with vertical or lateral configurations is illustrated in Fig. 3.

2
(a) (b)

Cathode Cathode

Anode

Perovskite
Substrate
Anode

Fig. 3 PV device architectures: (a) vertical and (b) lateral.

The photovoltage and photocurrent may depend on the BTO nanoparticle size, nanoparticle

concentration, magnitude of nanoparticle electric polarization, distance between the two opposite

electrodes, light intensity, electrical conductivity of the perovskite, crystallographic orientation,

perovskite/electrode interface.

The standard synthesis method of preparing methylammonium lead halide (CH3NH3PbI3-xClx)

perovskite precursor solution is mixing the powder of methylammonium iodide (MAI) with lead

chloride (PbCl2) at the 3: 1 molar ratio in Dimethylformamide (DMF). But MAI is commercially

not available and one must synthesize it first, which is a tedious process involving multiple steps.

Also, the chemicals being used to synthesize MAI (hydroiodic acid and methylamine) are quite

expensive. Therefore, this work was also driven by searching for a low cost, convenient new

synthesis method.

The basic research in the growth of the perovskite absorber and the use of an internal electric

field produced by ferroelectric nanoparticles those are embedded in the absorber is conducted

with the following research objectives.

1.1. Project Objectives

Three main project objectives.

3
(1). Introducing a novel, low-cost, single step, and very convenient method to synthesis

methylammonium lead halide perovskite solar absorbers.

(2). Fabrication of a ferroelectric BTO nanoparticles embedded methylammonium lead halide

perovskite photovoltaic device by the low-pressure spray deposition technique described in

section 2.4.

(3). Developing a simple computational model to guide the composite structure to optimize

optical absorption and the built-in internal electric field.

1.2. Research Plan

The project consists of four main phases towards the completion of above mentioned objectives.

Phase I: Precursor synthesis, deposition and characterization of crystalline methylammonium

lead halide perovskite solar absorber using PbI2 and CH3NH3Cl as starting precursors. The

research will include;

 Finding the optimum growth conditions; substrate temperature, precursor concentration.

 Structural characterization of the films through the use of SEM, EDS, and XRD.

 Optical characterization of the resulting thin films by UV-vis-NIR spectroscopy.

Phase II: Deposition and characterization of ferroelectric polymer PVDF: DMF using the low-

pressure spray deposition technique. The research will include;

 Determining the optimum growth conditions; PVDF concentration in DMF, substrate

temperature, film thickness.

 Structural characterization of the film morphology through the use of SEM, and XRD.

 Optical characterization of the resulting thin films by FTIR spectroscopy.


4
Phase III: Deposition and characterization of ferroelectric nanocomposite PVDF: BTO: DMF

using the low-pressure spray deposition technique. The research will include;

 Determining the optimum growth conditions; PVDF: BTO concentration in DMF,

substrate temperature, film thickness.

 Structural characterization of the PVDF: BTO structure through the use of SEM, and

XRD.

 Ferroelectric polarization measurements of the structure using Ferroelectric Tester.

 Investigate the effect of nanoparticle concentration, growth temperature, film thickness

on the ferroelectric properties of the polycrystalline film structure with poling and

without poling.

Phase IV: Deposition and characterization of ITO/CH3NH3PbI3-xClx:: BTO/Al structure using the

low-pressure spray deposition technique. The research will include;

 Determining the optimum growth conditions such as substrate temperature, film

thickness.

 Structural characterization of the capacitor through the use of SEM, and XRD.

 Optical characterization of the capacitor by FTIR and UV-vis spectroscopy.

 Ferroelectric polarization measurements with poling and without poling of the capacitor

using Ferroelectric Tester.

 Study the effect of nanoparticle concentration, growth temperature, film thickness on the

ferroelectric properties of the structure.

5
2. Background
Harnessing solar energy, which is abundant and sustainable, by means of solar technologies, in

particular photovoltaics, has attracted enormous interest in terms of research and development

for many years. Currently, the photovoltaic field is divided into three generations. The first

generation of solar cells refers to a single p-n junction of a crystalline Si (c-Si), exhibiting up to

25.6 % (Panasonic, Japan, 2014-04-11) power conversion efficiency (PCE). The maximum

thermodynamic efficiency for the conversion of unconcentrated solar irradiance into electrical

free energy in the radiative limit assuming detailed balance and a single threshold absorber was

calculated by Shockley and Queisser in 1961 to be about 31% [1]. The second generation of

solar cells included the use of amorphous-silicon, poly-crystalline-silicon or micro-crystalline-

silicon (a-Si, p-Si and mc-Si), cadmium telluride (CdTe-19.6 %) or copper (gallium) indium

selenide/sulfide. Despite the fact that fair to good performance has been obtained for these

devices, mass usage is limited either by their expensive fabrication steps, or by the low

abundance of the materials, or the toxic properties for some elements. The third generation of

solar cells, developed over the last decade, aims at conversion efficiencies beyond the S-Q limit,

stability, and use of environmentally benign and earth abundant materials with cost-effective

solution processing. They differ from the previous generations in a few aspects: (1) First

generation solar cells are configured as bulk materials that are subsequently cut into wafers and

treated in a “top-down” method of synthesis (silicon being the most prevalent bulk material).

Third generation solar cells are configured as thin-films (inorganic layers organic dyes and

organic polymers), deposited on supporting substrates or as nanocrystal quantum dots (QDs)

embedded in a supporting matrix in a “bottom-up” approach; (2) Third generation solar cells do

not necessarily rely only on a traditional single p-n junction configuration for the separation of

6
the photo-generated carriers. Instead, this generation includes the use of tandem cells, composed

of a stack of p-n junctions of low-dimensional semiconductor structures. Within the limit of an

infinite stack of a cascade with various bang gaps, covering a wide range of the solar spectrum,

the ultimate conversion efficiency at one sun intensity can increase to about 66%; (3) Third

generation solar cells can be configured as donor-acceptor (D-A) hetero-junctions, with

staggered electronic band alignment. The typical devices include die-sensitized solar cells

(DSSCs), organic solar cells (OPV) or polymer solar cells (PSC), quantum-dot (QD) solar cells

and perovskite solar cells.

2.1. Perovskite Solar Absorbers


Solar cells based on organic-inorganic perovskites as a light absorber made a major breakthrough

in recent years in the f i e l d o f photovoltaics. Generally, perovskite refers to calcium

titanium oxide mineral species composed of calcium titanate, with the chemical

formula of CaTiO3. The mineral was discovered in the Ural Mountains of Russia by Gustav

Rose in 1839 [2] and is named after the Russian mineralogist Lev Perovski. Later on, the

word ‘perovskite’ was borrowed to describe any material with the same type of crystal

structure as calcium titanium oxide, known as the perovskite structure. The general

chemical formula for pure perovskite compounds is AMX3, where ‘A’ and ‘M’ are

two cations of very different sizes, and X is an anion that binds to both. Halide perovskite

(AMX3) represents a big collateral series of perovskite family and it is reasonable to divide

them roughly into alkali-halide perovskite and organo-metal halide perovskite (Fig. 4).

7
Fig. 4 Crystalline systems of perovskites.

2.2. Organometal Halide Perovskites

Organometal halide perovskites structure and physical properties were first reported by Weber

in 1978 [3]. Interest in organometal halide perovskites continues to increase due to its appealing

features that include low material cost, ease of synthesis, suitable direct bandgap with high light

absorbance, band gap tunability, ultra-fast charge carrier generation [4], slow electron-hole

recombination rates [5], low temperature processability (~ 100 °C) [6], long electron and hole

diffusion lengths [7], microsecond-long balanced ambipolar charge carrier mobilities [5],

excellent thermal stability, and small exciton binding energy [19 - 50 meV]. In the case of

organolead halide perovskites, A is either methylammonium, ethylammonium or

formamidinium, and M is lead and X is either chloride, bromide or iodide.

2.2.1. Methylammonium Lead Halide Perovskites

Solar cells employing methylammonium lead halide perovskites as light harvesters were first

reported at the annual meeting of the International Electrochemical Society (IES) with a PCE of

2.2% in 2006 [8]. PCE of perovskite solar cells increased rapidly to 19.3 % [9] confirmed
8
efficiency by August 2014. The most widely studied methylammonium lead halides include

CH3NH3PbI3, CH3NH3PbI2Cl, CH3NH3PbI3-xClx, CH3NH3Pb(I1-xBrx)3, CH3NH3PbBr3-xClx and

CH3NH3PbBr3. Beyond the pure lead halide perovskites, CH3NH3PbI3-xClx, the mixed halide

perovskite is under intensive study for solar cell applications, which is derived from the pure

halide perovskite CH3NH3PbI3 via apical substitution of the iodine atom with a chlorine atom.

The incorporation of the chlorine atom adds a small perturbation to the equatorial iodine atoms

including a small contraction of the octahedron. The energy required for insertion of a chlorine

atom is 160 meV [10, 11].

Perovskites have wide direct band gaps which can be tuned either by changing the alkyl group,

or metal atom and halide. The bandgap (Eg) was calculated to be 1.57 eV (CH3NH3PbI3), 1.80

eV (CH3NH3PbBr3) and 2.34 eV (CH3NH3PbCl3) [12] for cubic structure, which agreed well

with the measured values (ca. 1.53 eV for CH3NH3PbI3, ca. 2.20 eV for CH3NH3PbBr3 and ca.

3.00 eV for CH3NH3PbCl3). When formamidinium HC(NH2)2+ or ethylammonium was

substituted for methylammonium in the A site of APbI3, Eg was changed to 1.47 eV for

HC(NH2)2+ [13] and to 2.2 eV for CH3CH2NH3+ [14] mainly due to changes in the chemical

interaction between Pb and I accompanying structural modifications.

The energy difference between the Fermi energy and the valence band maximum of

CH3NH3PbI3-xClx (2.2% chlorine ratio) was measured to be 1.1 eV, indicating a weak n-type or

intrinsic semiconductor [15] in contrast to CH3NH3PbI3, which was reported to be a p-type

semiconductor [16]. The work function of ITO is reported to be in the range of 4.3 - 4.9 eV, and

the work function of Al is in the range of 4.06 – 4.26 eV. The electron affinity of CH3NH3PbI3-

xClx is 4.15 eV.

9
Table 1. Properties of different lead halide perovskites.

The intrinsic carrier density in the lead halide perovskite was measured to be 109 cm-3, similar to

intrinsic semiconductors like silicon [13], however the electron mobility is much lower with 66

cm2V-1s-1. [13]. Recent measurements for the high-frequency charge mobility in CH3NH3PbI3-

xClx set a lower bound of 11.6 cm2V-1s-1and 8 cm2V-1s-1 for CH3NH3PbI3 [5]. The diffusion

length measured via photoluminescence decay experiments for electrons and holes in

CH3NH3PbI3 was measured to be around 0.1 μm [17], slightly larger for electrons than for holes.

In CH3NH3PbI3-xClx, electron diffusion length is about 1.9 µm and for holes it is about 1.2 µm

[20]. Because of higher charge carrier diffusion lengths, CH3NH3PbI3-xClx has a much larger

recombination lifetime than CH3NH3PbI3. The molar extinction coefficient of CH3NH3PbI3 is

about 1.5 × 104 (mol/L)-1 cm-1 at 550 nm, which is 2 - 3 times higher than those of organic dyes

widely adopted in solid‐state DSSCs [18]. This enables the perovskite to have complete light

absorption in films as thin as 660 nm to overcome the thickness limitations (of ~2 μm) of the

classic solid‐state DSSCs. The planar junction structured devices are more suitable if the diffusion

length of the charge carriers is longer than the depth of the light absorption, as demonstrated in

the CH3NH3PbI3-xClx based planar cells that have an efficiency of 15.4% [19].

10
2.2.2. Crystal Structure

The crystal structure of Methylammonium (MA) lead halide perovskites is depicted in Fig. 5.

Fig.5 Crystal structure of Methylammonium lead halide perovskites.

The smaller lead cations (Pb2+) are surrounded by six halide anions (Cl-, Br-, or I-) at the corners,

forming six‐fold coordinated octahedrons that extend to a three‐dimensional network by

connecting the corners. The perovskite structure is based on ionic bonds in a way that the total

charge is balanced, thus forming anionic corner-connected inorganic octahedra via covalent

bonds with organic cations in the cuboctahedral gap. The stability and the distortion of the

perovskite structure depend on the ratio of the (M-X) distance to the (A-X) distance, called the

Goldschmidt tolerance factor (t) [20],

t = (RM + RX)/√2(RA+RX)

where RA, RM, and RX are ionic radii of A, M and X, respectively for AMX3b .

Ideal perovskites have a cubic geometry if t = 1. Perovskites have a pseudocubic or distorted

cubic structure for stability when t deviates from 1. Such a distortion will affect the electronic,

optical, and dielectric properties of perovskite materials. An additional consideration for

perovskite formability is the octahedral factor (μ) [21], μ = RM/RX. In the case of the alkali metal

halide perovskite, formability is determined from the t-μ mapping where the perovskite is

11
stabilized for a tolerance factor ranging between 0.813 and 1.107 and an octahedral factor

ranging from 0.442 and 0.895 [21].

Table 2. Estimation of the A cation radii in AMX3.

The ionic radii of methylammonium lead halide (MAPbX3) perovskite are estimated based on

effective ionic radii, where the MA cation radii range from ∼158 picometers (pm) to ∼259 pm,

corresponding to one or two C−C (150 pm) or C−N (148 pm) bonds as in Table 2. Tolerance

factors for CH3NH3PbX3 are calculated at 0.85, and 0.83 for X = Cl, and I, respectively, based on

the radii of CH3NH3+ = 180 pm [22], Pb2+ = 119 pm, Cl- = 181 pm, and I- = 220 pm.

2.2.3. Temperature Dependent Structural Properties

Table 3. Temperature-dependent structural data of CH3NH3PbI3.

The temperature-dependent crystal structure of CH3NH3PbI3 perovskites is summarized in Table

3. The perovskite-phases are named with α, β, and γ and the non-perovskite δ-phase [23]. α is the

high-temperature phase for temperatures T > 327 K [24] and has a pseudocubical crystal

structure. Poglitsch et al. [25] reported that the position of the CH3NH3+ cation in CH3NH3PbI3

perovskite is only fixed in the orthorhombic phase at low temperature and could not be fixed at

12
room temperature due to the cubic symmetry requiring eight identical positions for the cations.

As a result, the tetrahedral-coordinated C and N atoms exhibit a disordered state inside the eight

tetrahedral of the cuboctahedron around the normal A positions (1/2, 1/2, 1/2) of the AMX3

perovskite. Therefore, the CH3NH3PbI3 perovskites have tetragonal β-phase (non-

centrosymmetric, space group I4cm [26, 13] with lattice parameters a = 8.855 Å and c = 12.659

Å) [27], with a slightly distorted PbI6 octahedral around the c axis at room temperature. The

exact parameters depend on molecular orientation [28]. In the α- and β-phase, the

methylammonium cations are disordered. Ferroelectric response like capacitance and non-ohmic

behavior, which might be responsible for hysteretic behavior in the current/voltage curves, could

be attributed to the reorientation of the methylammonium cations in an external field and the

resistance of the inorganic lead-iodide lattice [11, 13]. Due to the tilting of the octahedra during

the phase transition from α- to β-phase, the unit cell doubles its length, thus octuples its volume

and forms a super-cell [13].

For temperatures below 162 K, the perovskite undergoes a phase transition to orthorhombic γ -

phase (space group Pmc21 [29]), where the methylammonium cations are ordered. During the

phase transitions from α to β and γ, the octahedra are tilted and deform from the ideal octahedron

with respect to cubic phase. For decreasing temperatures, tilting and deformation effects increase

[23]. Phase transitions between the α-, β-, and γ -phases occur in the solid phase, while the

transition to the non-perovskite δ-phase happens in the presence of solvents [13]. A schematic

phase diagram of the different perovskite phases is shown in Fig.3. A reversible phase transition

was found from the tetragonal → cubic transition occurring at 57.3 °C upon heating and a cubic

→ tetragonal transition at 56°C upon cooling from a crystal chemistry study of CH3NH3PbI3

using a differential scanning calorimeter [30].

13
Fig. 6. Scheme ([006] view) of phase transitions of lead halide perovskite. The high-temperature
α-phase has two possible phase transitions, for dry crystals to the β- (T < 327 K) and γ -phase (T
< 162 K) or in the presence of, e.g., a solvent to the δ-phase for temperature below ∼360 K.
Stoumpos et al., Inorg. Chem. 52(15), 9019 (2013). Copyright 2014 American Chemical Society.

According to first-principles band calculations [31] and density functional theory (DFT)

calculations [32] of the CH3NH3PbI3 (X=Br, I) perovskites, if the spin orbital coupling is not

considered, the electronic bandgap (Eg) of CH3NH3PbI3 is 0.5 eV larger, than the experimental

data at room temperature. The reduction in Eg originates from the spin orbital coupling, in which

the valence band maximum created by the antibonding of Pb 6p - I 5p is nearly unchanged, but

the 6-fold nature of the conduction band minimum is primarily composed of Pb 6p - I 5s, a

significant change [33]. Therefore, the structural phase transition and spin orbital coupling could

affect the photovoltaic properties because the band structure is expected to be influenced by

structural changes.

14
2.2.4. Operational Principles of Perovskite Solar Cells

Light absorption, charge separation, charge transport, and charge collection are general solar cell

working processes. The excellent photovoltaic performance of perovskite solar cells is mainly

attributed to the excellent optical properties such as direct band structure and high light

extinction coefficient. In addition, photo-induced charges generate almost instantaneously after

photoexcitation and dissociate in 2 ps forming highly mobile charges (25 cm2V-1s-1) almost

balanced electron and hole mobilities remain very high up to the microsecond timescale. The

geminate recombination is, however, very slow, in tens of microseconds due to the crystalline

nature of the perovskite as well as low trap densities. Compared with other thin film

polycrystalline semiconductors, the perovskite absorber materials show a low non-radiative

recombination rate. The main recombination mechanism is free carrier recombination.

These characteristics provide the possibility of obtaining efficient photocurrent generation

efficiency for high short-circuit current density. The efficiency of charge extraction in

perovskite solar cells depends on the ratio between the charge separation and charge

recombination rates.

The most striking aspect of the perovskite technology is the high open-circuit voltage that the

cells can generate under full sun illumination. For the broad spectrum sunlight irradiating the

solar cell, photons with lower than band gap energy are not absorbed; all of the photoexcited

electrons from the above band gap excitations relax down to the conduction band, losing their

excess energy as heat. The maximum voltage that the solar cell can generate, the open-circuit

voltage, reflects the maximum energy that can be extracted from any absorbed photon, and the

difference between the potential of the lowest-energy absorbed photon that generates charge and

the open-circuit potential under full sunlight can be considered a simple measure of the

15
fundamental energy loss or “loss-in-potential” in a solar cell. Due to thermodynamic constraints,

this minimum loss in energy is on the order of 250-300 meV (depending on the band gap),

following the Shockley- Queisser treatment. The lowest energy absorbed photon that can

contribute to free carrier generation can be determined from the onset of the IPCE spectrum,

which for the CH3NH3PbI3-xClx perovskite is approximately 1.55 eV (800 nm), and the open-

circuit voltage of the best perovskite cells can be greater than 1.15 V, because the driving force

for hole extraction (0.2 V) is the only consideration. The photovoltage will be determined by the

difference between the Fermi energy of the perovskite and the HOMO energy of the HTM. In

Figure 7, the open-circuit voltage versus the optical band gap for the “best-in-class” of most

established and emerging solar technologies is shown.

Fig. 7 Open-circuit voltage versus optical bandgap.

The high mobility is also a critical factor for the efficient photovoltaic energy conversions, which

can generate a high fill factor (FF). The diffusion length of the charge carrier is an important

parameter for assessing the recombination probabilities in the materials. Long diffusion lengths

ensures the charges migrating to the interfaces before recombination. The interface engineering

is one good way to control charge transport and recombination.

16
2.2.5. Wannier Type Excitons and Origin of a High Open-Circuit Voltage

Since methylammonium lead halide perovskite contains both organic and inorganic parts,

electronic excitation can lead to Frenkel-type excitons with a high exciton binding energy and

small exciton Bohr radius and Wannier-type excitons with a low exciton binding energy and

large exciton Bohr radius. Exciton binding energy and exciton Bohr radius of CH3NH3PbI3 were

determined to be 19 - 50 meV and 30 Å, respectively [33]. Dielectric constant (ε) is estimated to

be 60. Since the Bohr radius is relatively large and the exciton binding energy is small, the

excitons from CH3NH3PbI3 will be Wannier-type. Frenkel excitons arise from organic

molecules, whereas Wannier excitons originate from the inorganic molecules, usually the lead

halide octahedral [34] part of methylammonium lead halide perovskites. Contrary to the 3-

dimensional methylammonium lead halide structures, low-dimensional structures have a higher

exciton binding energy of more than 200 meV [35-37]. With decreasing dimensionality to 1-

D and 0 - D, the exciton binding energies increase in accordance with the quantum

confinement effects [55,107]. CH3NH3PbI3-xClx indicate a more tightly bound nature of the

excitons resulting from the halogen substitution and the binding energy in mixed halide

perovskite is 35-75 meV [57].

In principle, the ionic compound has a permanent electric dipole moment and thus exhibits a

high dielectric constant. The dielectric constant of CH3NH3PbI3 perovskite is relatively high

because it is composed of a CH3NH3+ cation and PbX3- anion. Consequently, electric field

screening tends to reduce the Coulomb interaction between electrons and holes. Therefore,

Wannier type excitons generate through the illumination of light, because the exciton binding

energies are comparable to thermal energy (kBT ∼ 26 meV, where kB is the Boltzmann constant

and T is the temperature) at room temperature [7], and consequently the free charge carriers or

17
weakly bound excitons created building a high open-circuit voltage. The perovskites have a high

capacitance too owing to the linear relationship between the dielectric constant and capacitance.

Recently, Kim et al. reported that the CH3NH3PbI3 perovskite can accumulate charge carriers

within itself, like a capacitor, via impedance spectroscopy analysis [39]. This observation

indicates that free charge carriers exist within the perovskite film.

Since the methylammonium cations are polar by themselves and bonded via hydrogen bonds and

ionic effects, they can easily reorient within the lattice of inorganic octahedra. The reorientation

of the dipoles and lattice distortion due to the polar nature of the octahedra are the origin of some

effects of the lead halide perovskites. For instance, the absolute resistance shows a non-ohmic

hysteresis, which can be attributed to the reorientation of the methylammonium cations of the

perovskite under the influence of an external field [27]. These ferroelectric effects introduce

inductance and/or capacitance within the system, hence produce a large density of states within

the perovskites [39]. In addition to the polar nature of the lattice itself, the polarity of the organic

cations has a number of important effects on solar cell performance. The electronic polarization

of CH3NH3PbI3 and HC(NH2)2PbI3 is comparable to the polarization of materials, which show

above-bandgap voltages [40]. As reported by Choi et al. [41], a large fraction of the perovskite

consists of small crystallites, which could form ferroelectric domains [40]. Charge carriers in a

ferroelectric material like perovskite can diffuse via grain boundaries to the electrodes of the

device. In a multi-domain perovskite crystal, electrons would move along the minima of the

electric field and holes along the maxima, thus reducing recombination due to spatial separation

of the charge carriers [40]. Using Berry phase calculations, Liu et al. [53] found that the

polarization, aligning with the molecular dipoles, is mostly along the x axis (P x) with a value of

0.12 Cm-2 for MAPbI3. The hydrogen bond between the -NH3 group and the halogen atom is

18
critical for the bulk polarization. The dipoles of the MA cations are not perfectly aligned along

the x axis, the unit cell has a small component of polarization (∼0.04 Cm-2) along the z direction

(Pz). Their results confirmed that ferroelectric domains in hybrid halide perovskites help the

separation of photoexcited electrons and holes, and reduce recombination through the

segregation of charge carriers. Both 180 and 90° charged domain walls can serve as segregated

channels for the motions of charge carriers.

2.3. Ferroelectric Nanocomposites

Credit for discovering ferroelectricity goes to Joseph Valasek for his systematic

study of the magnetic properties of ferromagnetics and the dielectric properties of

Rochelle salt, which he presented at the annual meeting of the American Physical

Society in Washington on 23rd April 1920 [42,44]. In ferroelectrics, there must be a

bistable spontaneous electrical polarization that can be reversed, repeatedly, by an opposing

electric field to produce a polarization hysteresis [43, 44]. Reversibility of the spontaneous

polarization is an essential requirement to be a ferroelectric material.

Fig. 8 Typical hysteresis curve of a ferroelectric material.

Spontaneous polarization occurs in ferroelectric materials whose crystal structure exhibits

electrical order without the help of an external electric field. It is believed that the driving force

of spontaneous process is a thermodynamic free energy that leads the system into a lower

energy state, which is thermodynamically stable. Below the Curie temperature these materials
19
exhibit spontaneous polarization Ps and display hysteresis effects of polarization versus electric

field E.

The important characteristics of a polarization loop are, maximum polarization Pm, polarization

value at zero field the remnant polarization Pr, and the coercive field Ec which is the electric field

value when the polarization is zero. Above the Curie temperature, spontaneous polarization

disappears and the polarization response becomes linear - this state is called a paraelectric state.

Among 32 classes of crystals based on the degree of symmetry, only ten noncentrosymmetric

crystals exhibit spontaneous electrical polarization, that is, they have a permanent dipole

moment per unit volume in the direction of polar axis [45]. Ferroelectrics have recently attracted

attention as a candidate class of materials for use in photovoltaic devices [46-52].

Size and dimensionality plays a critical role in determining the ferroelectric

characteristics of a material at the nanoscale, due to the different ways in which dipoles

align in ferroelectric crystals. The scaling of a ferroelectric into ‘nano’ dimensions results

in an increase in the materials surface area, at which point surface charges play a

dominant role in determining the polarization of the nanomaterial. I n 1979, Kretschmer

and Binder highlighted the influence of surface effects on the spontaneous polarization

of ferroelectric thin films [109]. Nanostructuring can generate highly dense arrays of

isolated ferroelectric nanostructures and domains [110-112]. Since ferroelectricity is

a cooperative phenomenon caused by the arrangement of charge dipoles within a

crystal structure, an increase in surface area by nanostructuring could trigger immense

changes in the long- and short-range ordering of dipoles such as depolarization field

inside a material. These changes could alter some of the ferroelectric functionality, such

as phase transition temperature or Curie temperature (TC), domain dynamics,

20
dielectric constant, coercive field, or spontaneous polarization, piezoelectric response,

etc., at the nano scale [56-64]. In short, low dimensional ferroelectrics show marked

deviations in their properties compared to their bulk ferroelectric counterparts, mostly

due to the great enhancement of surface area. Since the surface characteristics of

nanostructures are morphology and size dependent, the type of nanonostructuring used

in ferroelectrics must be based on the dependence of some parameter related to the

ferroelectric functionality under consideration. This parameter could be crystallinity,

alignment, ordering or even surface modification of the ferroelectric nanostructures

under consideration [65].

There are two issues that must be carefully addressed when processing ferroelectric

nanocomposites (hybrid ceramic-polymeric composites). That is they have to be processed in a

way that enhances its crystallinity and favors the growing of the polar phase. Recently, PVDF

has been intensively studied by many authors as a polymer matrix for ceramic nanopowders such

as BaTiO3 [66-69], PbTiO3 [70], CaCO3 [71], and Pb(Zr0.5TiO0.5)O3 [68] because they combine

the excellent ferroelectric properties of ceramics with the flexible mechanical and ferroelectric

properties of the polymer. Lijie et al. reported the fabrication and characterization of

BaTiO3/PVDF nanocomposites via the sol–gel method, in which nanosized BaTiO3 particles

with an average size of 50–100nm were grown in situ in the PVDF matrix [67]. BaTiO3 is a

ferroelectric ceramic widely used in capacitors and ultrasonic transducers. It was observed that

the relative dielectric constant of nanocomposites increased in the frequency range of 5×104 to

3×106 Hz with increasing weight fraction of nanosized ceramic in the polymer matrix. According

to Mao et.al, the remanent polarization in BaTiO3/PVDF composites increases with the BaTiO3

particle size increasing from 25 to 500 nm. The dielectric constant of the composite with BaTiO3

21
nanoparticles over 250 nm remains almost constant, while it increases with decreasing particle

sizes and reaches a maximum value at a BaTiO3 particle size around 80 - 100 nm. The dielectric

constant decreases with further decrease in particle sizes down to 50 nm and then increases from

50 to 25 nm, showing a minimum at 50 nm [72].

2.3.1. Ferroelectric Photovoltaic Effect

Ferroelectric materials have been of great interest for photovoltaic applications due to its

switchable rectifying behavior, polarization-dependent photovoltages, the efficient electron-hole

pair separation at ferroelectric domain walls, and ultrafast carrier dynamics under above band-

gap femtosecond optical excitation [73-75]. The ferroelectric photovoltaic (FEPV) effect was

discovered about half a century ago in a variety of ferroelectric materials [76, 77]. The

conversion process of light energy to electrical energy in photovoltaic devices relies on some

form of built-in asymmetry that leads to the separation of electrons and holes. The fundamental

physics behind this effect (for example, in silicon-based cells) is charge separation using the

potential developed at a p-n junction, or heterojunction [78-80]. On the other hand, ferroelectrics

respond to light illumination to generate the steady-state photovoltage and photocurrent because

of the separation of photo-induced charge carriers by the internal polarization electric field. The

magnitude of this photovoltage is directly proportional to the crystal length in the polarization

direction. In the junction-based photovoltaic effect, the internal electric field only exists in a very

thin depletion layer at the interface. Without the internal field in the bulk material region, the

photogenerated charge carriers swept out of the depletion region have to diffuse to the cathode or

anode. Thus the charge transportation is often limited by diffusion in the junction-based

photovoltaics. The open circuit photovoltage cannot exceed the energy barrier height of the

junction, which is usually lower than 1 V. In contrast, for the ferroelectric bulk photovoltaic

22
effect, the remnant polarization and the polarization-induced internal electric field exist over the

whole bulk region of the ferroelectric rather than a thin interfacial layer. In this case, the charge

transportation is not limited by diffusion, and the output photovoltage is not limited by any

energy barrier. For example, by poling a ferroelectric thin film in the plane of the surface to

break the constraint of the small thickness, a large photovoltage of 7.0 V was demonstrated in

ferroelectric thin films [81]. Previous studies have revealed that the ferroelectric film materials,

including BaTiO3 (BTO) and Pb(Zr, Ti)O3 (PZT) thin films, exhibit photovoltaic effects in the

ultraviolet (UV) region resulting from the relatively wide band gap (typically larger than 3.3 eV)

[82-84].

2.3.2. Bismuth Ferrite (BiFeO3 (BFO))

BiFeO3 (BFO) thin film (Eg ~ 2.67 eV), which has high sensitivity and fast response to the

visible light [85, 86], offers a promising opportunity for solar photovoltaic applications. Bulk

bismuth ferrite can be described as a rhombohedrally distorted ferroelectric perovskite with the

space group R3c. The lattice parameters of the rhombohedral unit cell are a = 5.63 Å and α =

59.4˚. In such a distorted structure, the R3c symmetry permits the development of spontaneous

polarization (Ps) along the pseudocubic (111) direction and (100) direction. Spontaneous

polarizations in a single crystal have been measured at 80 K to be 3.5 μC/cm 2 along a (100)

direction and 6.1 μC/cm2 along a (111) direction [87]. The enhanced polarization of about 100

μC/cm2 was observed in the thin films fabricated by pulsed laser deposition [88, 89] as well as

chemical solution deposition [90]. Remnant polarization (>60 µC/cm2), and Curie temperature

(~1103 K) [88, 91] of BFO are relatively high.

Remarkable FEPV effects have already been reported in single crystalline BFO bulk, BFO

ceramic and epitaxial BFO thin films [92-97]. A very attractive feature of ferroelectric

23
photovoltaic materials is their tunable photovoltaic output with the external stimuli [98].

Recently Ji et al. [96] found that photocurrent direction can be reversed by applying an external

electric field to switch the polarization direction of the epitaxial BFO film, which indicated that

the ferroelectric polarization played a dominant role in the observed photovoltaic effect. And

polycrystalline BFO (poly-BFO) films may be more suitable for the practical applications due to

high cost and complicated fabrication process of epitaxial BFO thin films. To study the

ferroelectric photovoltaic effect based on polycrystalline films, preparation of high-quality

polycrystalline BFO films with low leakage and high remnant polarization is essential. Doping of

foreign atoms into BFO is effective for the control of leakage current. The leakage current

density in ferroelectrics corresponds to the dark current density or the dark conductivity in

photovoltaics.

2.3.3. Origin of the large photovoltage in FEPV devices

The origin of the FEPV effect is controversial and in order to explain the ultra-high photovoltage

output, several models have been proposed in early years, including the shift current model and

the nonlinear dielectric model [102]. The common characteristic of these theories is that the

photovoltage is generated in the bulk of the ferroelectric crystals, hence named as the bulk

photo-voltage effect. A recent theory gives an alternative explanation on the origin of the FEPV

effect using a series of domain walls in tandem with each other outputting a small photovoltage

[47]. Other effects related to the ferroelectric/electrode interface, e.g. Schottky effect and

screening effect [103-106], are also believed to generate or influence the photovoltage output in

ferroelectric thin films. These theories are related to the domain wall interface or the

FE/electrode interface. Yang et al.[47] proposed the Domain wall theory. They studied the FEPV

effect on the BFO film with ordered domain strips and lateral device configuration (Fig. 9). They

24
observed that the photovoltage in the BiFeO3 films increased linearly with the total number of

domain walls along the net polarization direction (perpendicular to the domain walls, Fig. 9a)

[47].

Fig. 9 Parallel and perpendicular configurations of photovoltage measurements of BiFeO3 films.

The photovoltaic effect vanished along the direction perpendicular to the net polarization

direction (Fig. 9b). The intrinsic potential drop at domain walls (~10 mV), arising from the

component of the polarization perpendicular to the domain wall, induces a huge electric field of

~ 5 ×106 Vm-1 in the narrow domain wall, which was suggested to be the driving force for the

dissociation of the photogenerated exciton. The illuminated domain walls act as nanoscale

photovoltage generators connected in series, wherein the generated photo-current is continuous

and the photogenerated voltage accumulates along the net polarization direction. This proposed

mechanism is analogous to the concept of tandem solar cells, where the output voltage is the sum

of the photovoltage of each sub-cell. The evolution of the photo-response of the BFO films as a

function of incident light intensity is shown in Fig. 10.

Fig. 10 Photovoltage vs light intensity of BFO films.

25
Above light powers of ~ 15 mW (corresponding to a power density of ~285 mW/cm2) VOC was

found to saturate to a value of ~ 8 mV / domain wall.

Some of the present studies are based on the transparent electrode/poly-BFO film/Pt (substrate)

structure, and the incoming sunlight is incident from the top electrodes such as indium tin oxide

(ITO) [85, 99], grapheme [86] and carbon nanotube (CNT) [100]. Furthermore, Zang et al. [100]

discussed the temperature dependence of the photovoltaic output in the CNT/BFO/Pt

heterojunction. As for the top electrode/poly-BFO film/FTO (substrate) capacitor, Guo and Liu et

al. made some efforts to enhance the photovoltaic performance by comparing different top

electrodes including Au, ITO and aluminum-doped zinc oxide (AZO) [101].

2.3.4. PV properties of Pt/ BiFeO3 (BFO) thin film/FTO capacitor

Zhou et al.[146] studied the photovoltaic properties of the Pt/ BiFeO3 (BFO) thin film/fluorine-

doped tin oxide capacitor. They compared two BFO films and as compared to the BFO film with

random orientations (B), the BFO film with a strong preferred orientation (A) exhibited larger

photovoltaic output as a result of its larger spontaneous polarization in the unpoled state.

The nonzero photovoltage VOC in Fig. 8 indicates that there is an internal electric field (Ein) as a

driving force to separate the photoinduced electron-hole carriers in the capacitor. For the

ferroelectric materials, the Ein comes mainly from the polarization field in the film and built-in

electric field due to the Schottky barriers at the electrode/film interfaces [82, 118]. The unpoled

BFO film is self-polarized and it can readily function as a photovoltaic cell [94]. So, as compared

to BFO film B, the larger photovoltage VOC of film A is a consequence of the larger spontaneous

polarization in the unpoled film A.

26
Fig. 11 Dark and light illumination J–V characteristics for BFO films.

To probe the role of external electric field polarization on the photovoltaic output, the samples

were polarized along the two opposite directions, namely the switchable polarization from the

positively to negatively poled state was carried out by applying electric field of +370 and -370

kV/cm to the Pt electrode, respectively. The photovoltaic output variations with external electric

field polarization are presented in Fig. 12.

Fig. 12 Illuminated J-V characteristics of BFO films in different polarization states.

Tthe external electric field poling has noticeable influence on the photovoltage in the two BFO

films, and the net change of the photovoltage after positive and negative poling is larger in the

27
BFO film A than in the BFO film B. The net change in the photovoltaic output after poling is

largely depend on the structure and ferroelectric polarization [147].

Fig. 13 Ferroelectric polarization - electric field hysteresis loops measured at a maximum

applied electric field of 400 kV/cm and 10 kHz.

The larger net change of the photovoltage in the BFO film A is may be due to its improved

crystalline structure originated from the strong preferred orientation and hence larger remnant

polarization. The P-E measurements indicate that the BFO film A possesses larger remnant

polarization than the film B at the same applied electric field and frequency.

The Pt/BFO film/FTO structure can be regarded as the semiconducting BFO [138] connected

with two back-to-back Schottky diodes in series. The ideal Schottky barrier at a

metal/semiconductor interface is dependent on the difference of the metal work function and the

semiconductor electron affinity. The work function of FTO is reported to be in the range of 4.7 -

4.9 eV [139-141], and the work function of Pt is 5.3 eV. The electron affinity of BFO is 3.3 eV

[142,143]. Therefore, the ideal Schottky barrier height at the top interface (φtop) is higher than

that at the bottom interface (φbottom) in the BFO capacitor. Ebi-top and Ebi-bottom represent two built-

in electric fields existing in the top Pt/BFO and bottom BFO/ FTO interfaces, respectively. In

comparison with the unpoled state (Fig. 12), the photovoltage for positively poled state

28
(downward polarized state) decreases significantly, and the one for negatively poled state

(upward polarized state) exceeds the initial value for the unpoled state. It suggests that the self-

polarization field (perpendicular to the film surface) is toward the bottom FTO electrode. For the

unpoled BFO film, the resultant direction of internal electric field is from Pt to FTO electrode

because the photocurrent flows out of the FTO electrode. Figure 14 a shows the energy band

schematic diagram of the Pt/BFO film/FTO capacitor, and the Esp is the spontaneous polarization

field.

In the BFO film based capacitor, the oxygen vacancies as donor impurities in BFO are easy to

release electrons to conduction band and become positively charged under the high external

polarization electric field, and moreover the depletion regions or Schottky barriers at both bottom

and top electrodes can be modulated by the redistribution of carriers in the polarized BFO thin

film [144].

In the positively poled state, the upward depolarization field (Edp) can drive the released

electrons to neutralize the positive bound charges at the bottom BFO/FTO interface and a

neighboring narrow electron region in BFO film can be formed near the bottom interface with

positive bound charges, resulting in the downward-bending of the band and form of the ohmic

contact at this interface; meanwhile, near the top Pt/BFO interface with negative bound charges,

oxygen vacancies become positive charged and the depletion region becomes wide, accompanied

with upward band bending and an enhanced Schottky-like barrier [144, 145]. In summary, the

directions of depolarization field and built-in electric field due to the Schottky barrier are same

and all towards the top Pt electrode in the positively poled BFO film. Figure 14 b and c show the

modulated band structure diagrams for the positively and negatively poled states, respectively.

29
Fig. 14 Energy band schematic diagrams of Pt/BFO film/FTO capacitor in the unpoled state (a),

in the positively poled state (b) and in the negatively poled state (c) (not to scale).

2.3.5. Barium Titanate (BaTiO3 (BTO)) Nanoparticles


Ferroelectricity in barium titanate (BaTiO3) was discovered in early 1940s. BTO has the

perovskite crystal structure, which has the general formula ABO3, where A represents a divalent

metal ion (barium) and B represents tetravalent metal ions (titanium in this case). Above the

Curie temperature (TC), the crystal has a cubic symmetry, a centrosymmetric microstructure

where the positive and negative charges coincide. . It transforms to the tetragonal (T) phase, the

orthorhombic (O) phase, and the rhombohedral (R) phase, which are all ferroelectric, when the

temperature decreases to 130 °C (Tc), 5 °C, and -75 °C, respectively [108].

A consequence of nanostructuring ferroelectric materials is the appearance of a critical

size limit, below which spontaneous polarization cannot be sustained in a ferroelectric


30
material [54, 60]. Pithan et al. [128] used the Raman spectroscopy method to confirm the

presence of tetragonality in nanocrystalline BaTiO3 powders (10 nm) and ceramics (35 nm).

Zhao et al. [129] evaluated the critical size for disappearance of BaTiO3 ceramics ferroelectricity

to be 10-30 nm. The grain size was further reduced to 5nm in 2011 [130]. Yashima et al. [116]

indicated the presence of both bands of the tetragonal and hexagonal phases in the 40-nm-sized

particles by Raman spectrum. Core-shell structures were found in nano BaTiO3 grains using

TEM technique [123,117]. Xiao et al. [119] found the coexistence of ferroelectric tetragonal and

orthorhombic phases in 30 nm BaTiO3 ceramics at room temperature. For the core-shell structure

in nano BaTiO3 ceramics, Huang et al. [148] suggested that the presence of hexagonal symmetry

be attributed to the coexistence of cubic/tetragonal phases at nanometer scale. Wada et al. [120-

122] put forward a two-phase model consisting of tetragonal and cubic forms. Zhao et al. [129]

admitted the existence of a paraelectric or quasi-paraelectric shell on the surface of the - BaTiO3

grains near the grain boundary. Hoshina et al. [126] believed that BaTiO3 nanoparticles had the

composite structure consisting of inner tetragonal core, gradient lattice strain layer, and the

surface cubic layer which induces phase transition in the nanocrystalline ferroelectric BaTiO3.

According to Li et al. [114], each BaTiO3 nanoparticle is surrounded by several different

surfaces with complicated structures. Atomic shifts induced by surface relaxation will reduce the

symmetry of the shell structure of the nanoparticle. The thickness of the relaxed shell is about

1nm, accounting for a sharp increasing share of the particle as the particle size decreases to only

several nanometers. Surface relaxation is the origin of the coexistence of various phases in

BaTiO3 nanoparticles.

31
Fig. 15 Curie temperature dependence on BTO particle size.

The multi-phase coexistence is found in BaTiO3 nanoparticles at the temperature range of 25 °C

to 600 °C, implying that ferroelectric phases can remain at 600 °C and the Tc significantly

increases from 130 °C to at least 600 °C. Temperature-size phase diagrams were proposed to

reveal the dual effect of size and temperature in BaTiO3. In the phase diagrams, the point on the

phase boundary indicates the phase transition temperature with a given size and the critical size

of phase transition at a given temperature as well. Fig. 15 shows the temperature-particle size

phase diagram for BaTiO3 particles. Particles larger than 1 µm are treated as bulk BaTiO3, and

have the normal phase structures as shown in the right half part of the phase diagram. The results

of high resolution synchrotron x-ray diffraction [130, 115] are summarized in the left half part of

the diagram, according to which the phase boundaries have been drawn. As shown in Fig. 15

experimental data have been filled into the temperature-particle size phase diagram with red

markers. All the experimental particles are composed of cubic, tetragonal, orthorhombic, and

rhombohedral phases.

32
2.4. Low-Pressure Spray Deposition

Spin coating, doctor blading, slot-die coating, gravure coating, knife-over-edge coating, off-set

coating, spray coating, ink jet printing, pad printing and screen printing are a few techniques

being used to fabricate solution-processed solar cells. The characteristics of fabricated films

such as film morphology, density, crystallinity, roughness, and structure depend on the technique

being used. Among aforementioned methods; film growth by spray coating is very attractive

since its capability of making large area coatings. The main step of spray coating is the

atomization of a liquid precursor solution to generate microdroplets. Transportation of

microdroplets to the substrate followed by droplet impact on a heated substrate, droplet

spreading, solvent evaporation, drying, solute adhesion and bonding to itself and to the substrate

leads to the formation of a thick coating. One-step precursor solution deposition [131], two-step

sequential deposition [132], dual-source vapor deposition [133] and vapor assisted solution

process [134] are the reported fabrication techniques for organometal lead halide perovskites.

Low-vacuum spray deposition is a simple technique that can be used to fabricate highly

crystalline methylammonium lead halide perovskite thin films and the ferroelectric phases of

PVDF/ BaTiO3 nanocomposites directly from solution. This technique can be easily adapted to

form ferroelectric nanocomposites when the nanoparticles are suspended in the polymer solution.

The schematic diagram of the experimental apparatus is shown in Fig. 16.

Films are usually deposited on glass substrates or ITO (Indium tin oxide) coated glass substrates.

First the precursor is nebulized into 1-5 µm aerosols using a nebulizer and injected into the low-

pressure deposition system using nitrogen as the carrier gas, since as produced particles lack

sufficient inertia. Since the aerosol-generating rate is independent of the flow rate, the transport

of the droplets to the substrate can be controlled without affecting the function of the nebulizer.

33
N2 gas Substrate
Holder

N2 gas

Nebulizer

to Vacuum
Precursor
pump
Solution

Fig. 16 Experimental setup for the low-pressure spray deposition process.

The substrate was mounted on a heating block to promote film growth. The spray chamber was

characterized; it was found that 6 cm distance from tip of the nozzle to the substrate is the ideal

distance and 550 Torr is the ideal pressure to get a good coating. When the droplets impinge

upon the heated substrate the subsequent film formation is dependent on the velocity of the drop,

rate of reaction and the rate of evaporation of the solvent. At high velocities the droplets will

flatten on the substrate leading to large particle sizes. On the other hand, if most of the solvent is

evaporated when the drop gets to the substrate, the solid particles will stick to the substrate to

form a crystallite smaller than the initial droplet. Therefore by adjusting the concentration of the

solvent, the size of the particles depositing on the substrate can be controlled.

Fig. 17 Low-pressure spray deposition system.

34
2.5. Current Project Accomplishments

Investigations for Phase I of the research plan have been completed. Optimum growth conditions

have been determined for successful deposition of highly crystalline CH3NH3PbI3-xClx perovskite

films by the low-pressure spray deposition technique. To the best of our knowledge this is the

first reported successful growth of CH3NH3PbI3-xClx films starting from MACl and PbI2

precursors. Structural and optical ccharacterization of the deposited films have also been

completed. These results are presented in Chapter 3.

Preliminary studies for Phase II, deposition and characterization of ferroelectric polymer PVDF:

DMF using the low-pressure spray deposition technique are currently in progress. Study of the

PVDF: DMF concentration, substrate temperature, and film thickness on the ferroelectric

properties of the deposited films are currently being investigated.

2.6. Project Challenges

The first challenge for the proposed project is to be able to nebulize the precursor solutions.

Utilization of 2.4 MHz ultrasonic nebulizers for the deposition of CH3NH3PbI3-xClx perovskite

films was very inefficient because Au coating of the vibrating crystal is not resistant to Cl of the

mixed halide. We were able to design and utilize a vibrating-mesh nebulizer to overcome this

challenge. Schematic diagram is available in Appendix A.1. PVDF: DMF precursor solution

does not nebulize at higher concentrations. And upper concentration limit for the nebulization is

1.0 wt % of PVDF powder.

Achieving the optimum deposition temperature of the hybrid polycrystalline structure

(ITO/CH3NH3PbI3-xClx: PVDF: BTO/Al) will be another challenge, because preservation of the

highly crystalline structure of the mixed halide perovskite is highly important. 120 °C is the

optimum growth temperature for the mixed halide perovskite. The Curie temperature of BTO

35
nanoparticles is in the range of 25-600 ºC. Identification of Curie temperature of BTO

nanoparticles for a given size is important.

The deposition of Al top contacts, in-situ will be another challenge, as we have not previously

attempted this in the laboratory. Another challenge would be to have control on uniform film

thickness. Thin films deposited by low-pressure spray deposition have a sharp thickness profile

depending on growth parameters. This could be an obstacle as non-uniformity in film thickness

could lead to inconsistent photovoltage measurements. One of the possibilities to overcome this

could be by rotating the substrate during deposition.

36
3. Results
3.1. Synthesis of CH3NH3PbI3-xClx Perovskites
The precursors used for the synthesis of CH3NH3PbI3-xClx were prepared by dissolving

CH3NH3Cl (or MACl) and PbI2 with the molar ratio of 3:1 in a glove box followed by 2 h

sonication. Six different precursors were prepared by dissolving appropriate number of moles of

MACl and PbI2 in 3 ml of DMF. Basically, 0.11 M, 0.22 M, 0.275 M, 0.33 M, 0.385 M, and 0.44

M concentrations of PbI2 and 0.33 M, 0.66 M, 0.825 M, 0.99 M, 1.155 M, and 1.32 M

concentrations of MACl were used for each precursor solution. Glass substrates were cleaned by

sonicating for 10 minutes in liquinox detergent, acetone, iso-propanol, and deionized (DI) water

respectively.

3.2. Deposition of CH3NH3PbI3-xClx Perovskite Films

The deposition system utilized in the experiments is described in Chapter 2.4 with a diagram of

the low-vacuum spray deposition system shown in Figure 1.16. The vacuum chamber was

evacuated by a roughing pump. The substrate heater is a stainless steel block internally heated by

a 600 watt halogen bulb. The substrate temperature is monitored by a K-type thermocouple with

the tip in thermal contact by conductive silver paint to the face of the heater block. The substrates

are located adjacent to the thermocouple tip and fixed to the heater block with conductive silver

paint to ensure thermal contact. The substrate temperature was recorded as the surface

temperature of the heater block. The substrate was mounted on a heating block that was placed 6

cm in front of the nozzle and the pressure of the deposition chamber was kept at 550 Torr. Films

deposited with precursors 1 to 6 were held at deposition temperature for additional 5 minutes to

evaporate any residual solvent and the effect of substrate temperature and precursor

concentration was studied.

37
3.3. Characterization of CH3NH3PbI3-xClx Perovskite Films

Powder X-ray diffraction (XRD) was performed by Bruker AXS D8 Focus X-ray diffractometer

with graphite monochromatized Cu Kα1 radiation (λ = 0.15406 nm). The accelerating voltage

was set to 40 kV with 40 mA flux at a scanning rate of 5°/min in the 2θ range from 12° to 60°.

SEM was carried out using JEOL JSM-6390LV Scanning Electron Microscope. The SEM spot

size and acceleration voltage was kept at 30 and 20 kV respectively. The magnification was

varied from 3000x to 30,000x. It is also equipped with an energy dispersive spectroscopy (EDS)

detector from Oxford Instruments INCA x-sight for compositional analysis. A Perkin-Elmer

Lambda 950 UV-vis-NIR spectrometer was used to measure the absorption spectra of

CH3NH3PbI3-xClx perovskite films. A Veeco Dektak 3030ST profilometer was used to measure

the thickness of CH3NH3PbI3-xClx perovskite films.

3.3.1. Structural Characterization

X-ray diffraction patterns of CH3NH3PbI3-xClx perovskite films fabricated from different

precursor concentrations of PbI2 and MACl are shown in Fig 18. After a preliminary study, 120

°C was used as the deposition temperature. The crystal structure evolution of the CH3NH3PbI3-

xClx films on the concentration dependence was examined by the XRD patterns. When PbI2 and

MACl concentrations are 0.11 M and 0.33 M respectively, the film exhibits a strong XRD peak

near 12.65° which corresponds to a low-level impurity or excess PbI2. It is interesting to note

that sample (d) produced the strongest and narrowest CH3NH3PbI3-xClx (110) peak near 140 and

the strongest absorption (Fig. 22). This may suggest the sample (d) condition is the optimum

condition to grow highly crystalline CH3NH3PbI3-xClx perovskite films. The perovskite sample

(d) shows diffraction peaks at 14.20°, 28.58°, 43.27°, and 58.88°, which can be assigned to the

38
(110), (220), (330) and (440) planes, respectively, of a tetragonal perovskite structure with lattice

parameters a = b = 8.825Å, c = 11.24Å, for CH3NH3PbI3-xClx as reported elsewhere.

Fig. 18 X-ray diffraction patterns of fabricated CH3NH3PbI3-xClx perovskite films (a) 0.11 M &

0.33 M (b) 0.22 M & 0.66 M (c) 0.275 M & 0.825 M (d) 0.33M & 0.99 M (e) 0.385 M & 1.155 M

(f) 0.44 M & 1.32 M PbI2 and MACl concentrations respectively.

The extremely narrow diffraction peaks suggest that the films have long-range crystalline

domains. When PbI2 and MACl concentrations are above 0.33 M and 0.99 M respectively, the

films exhibit different peak ratios, implying different orientation of grains but having the same

tetragonal crystal structure. This confirms that this low pressure spray technique is capable of

growing strongly oriented semi-single crystalline perovskite films. Even though, Colella et

al.[136] and Zhao and Zhu [137] reported adding methylammonium chloride to lead iodide

39
resulted in the segregation of a CH3NH3PbCl3 phase indicating low solubility of chlorine in the

iodine derivative, the resulted optimum condition diffraction pattern shows no signs of an excess

PbI2 or CH3NH3PbCl3 phase. This result confirms the complete conversion of MACl: PbI2

precursor material into CH3NH3PbI3-xClx and the possibility of the formation of pure, highly

crystalline CH3NH3PbI3-xClx perovskite films starting from methylammonium chloride and lead

iodide.

To study the effect of post-film deposition annealing, we increased the annealing time. Here, the

perovskite films were spray deposited using an ultra- sonic nebulized precursor solution

dissolved in DMF and then the films were annealed at 120 °C for 45 minutes. Similar XRD

patterns and absorption data to the 5 minutes annealing were observed. Since the ultra- sonic

nebulizer forms microdroplets of the precursor solution, there is a tiny volume of the solvent to

be evaporated from the heated glass substrate when the microdroplets deposit on the substrate.

Therefore, 5 minutes annealing is sufficient to get the optimum crystallization for the formation

of pure CH3NH3PbI3-xClx with good crystallinity and strong absorption.

Fig. 19 shows the typical scanning electron microscopy (SEM) images of top views of the spray

deposited perovskite films prepared at 120 Celsius and annealed for 5 minutes from precursor

solutions with different concentrations of PbI2 and MACl. At the magnification of 3000 x (see

the scale bars in Fig. 18 a), the CH3NH3PbI3-xClx perovskite films exhibit clustered-domain

patterns on the preheated glass substrate for sample (a).As shown in Fig. 20 a, the lengths of the

domain regions are approximately 5-10 μm. The clustered-domain pattern disappeared at higher

concentrations. The perovskite particle size increases as the concentration of precursor solution

increases.

40
(a) (b) (c)

(d) (e) (f)

Fig. 19 a Top view SEM images of fabricated CH3NH3PbI3-xClx perovskite films on glass
substrates at different concentrations ((a) 0.11 M & 0.33 M (b) 0.22 M & 0.66 M (c) 0.275 M &
0.825 M (d) 0.33M & 0.99 M (e) 0.385 M & 1.155 M (f) 0.44 M & 1.32 M PbI2 and MACl
concentrations respectively.

(a) (b) (c)

(d) (e) (f)

Fig. 19 b Top view SEM images of fabricated CH3NH3PbI3-xClx perovskite films on glass

substrates at different concentrations ((a) 0.11 M & 0.33 M (b) 0.22 M & 0.66 M (c) 0.275 M &

0.825 M (d) 0.33M & 0.99 M (e) 0.385 M & 1.155 M (f) 0.44 M & 1.32 M PbI2 and MACl

concentrations respectively.

41
The CH3NH3PbI3-xClx perovskite particle sizes are approximately 0.5 µm – 2 µm in diameter and

the CH3NH3PbI3-xClx perovskite films exhibit densely interconnected crystalline strips on the

glass substrate as seen at a magnification of 30000 x. It is worth noting that the SEM images of

CH3NH3PbI3-xClx perovskite films in Fig 19 are not the ideal film morphology, but the spray

deposited films on preheated glass substrate still has a smooth and reflective topography when

viewed by the naked eye.

Fig. 20 X-ray diffraction patterns CH3NH3PbI3-xClx perovskite films at different deposition

temperatures.

The crystal structure evolution with the substrate temperature was examined by the perovskite

XRD patterns. For that, we fabricated several samples at different deposition temperatures

ranging from 60 0C to 160 0C. Here each sample was prepared by 0.33 M PbI2 and 0.99 M MACl

dissolved in DMF followed by sonicating 2 hours. X-ray diffraction patterns are shown in Fig.

20 and surface morphology of those films are shown in Fig. 21. At low deposition temperatures

42
(80 °C), perovskite film has a different crystal structure indicating incomplete crystallization. The

surface looked rough and cracked to the naked eye. From the 100 °C to 140 °C deposition

temperatures, all the deposited perovskite films show the same XRD patterns with tetragonal

crystal structure with strongly oriented crystals. It is interesting to note that the strongest

CH3NH3PbI3-xClx (110) peak produced at 120 °C, and also became narrower indicating enhanced

crystallinity and a smooth and shiny film surface. At 160 ° C the perovskite film is decomposed

leaving a yellowish color mainly due to lead iodide. This study confirms that the optimum

deposition temperature is 120 °C for growing perovskite films via low pressure spray deposition.

80 °C 100 °C 110 °C

120 °C 130 °C 140 °C

Fig. 21 Top view of SEM images of fabricated CH3NH3PbI3-xClx perovskite films on glass

substrates at different deposition temperatures.

Typical Scanning electron microscopy(SEM) images at different temperatures are shown in Fig.

21. At low deposition temperatures (80 °C), the precursors produced unique rod-like structures.

Nice individual CH3NH3PbI3-xClx perovskite particles can be seen in the film fabricated at 120

°C which was determined to be the optimum deposition temperature. Above the optimum

43
temperature, perovskite particles looked fused together due to high temperature. Overall

appearances of all the samples deposited at 100 °C, 110 °C, and 140 °C are similar.

3.3.2. Optical Characterization

Fig. 22 Linear absorption spectra of CH3NH3PbI3-xClx perovskite films for 6 precursors.

The absorption spectra of as fabricated CH3NH3PbI3-xClx perovskite films are shown in Fig. 22.

The absorbance increases slightly with decreasing wavelength from approximately 750 to 400

nm. There is a sharp decrease in the absorption when the wavelength approaches the bandgap

near 800 nm. Here, we did a quantitative study of each absorption graph to determine the best

44
precursor sample. The ratio between total absorption within specified wavelength range (from

400 nm to 850 nm) to the absorption at the band edge (800 nm) as a percentage for precursors (a)

to (f) are 36.5%, 48.1%, 66%, 100%, 65.4%, and 82.8% respectively. Clearly, perovskite sample

(d) has the optimum absorption hence precursor (d) is the optimum condition for the synthesis of

CH3NH3PbI3-xClx perovskite. The optical bandgap (Eg) of CH3NH3PbI3-xClx can be determined

from the Tauc plot of the absorption spectrum. The Tauc plot follows the expression (αhν) = β

(hν – Eg)n, where α is the absorption coefficient, hν is the photon energy, Eg is the bandgap, and

n is 0.5 for a direct bandgap semiconductor. By extrapolating the linear region of the Tauc plot,

the bandgap is found to be 1.55 eV, which is consistent with reported bandgap for CH3NH3PbI3-

xClx.

The linear absorption spectra of fabricated CH3NH3PbI3-xClx perovskite films on glass substrates

at different deposition temperatures are shown in Fig. 23. The maximum absorption spectrum

was observed at 120 °C which is in agreement with XRD patterns.

To understand the effect of post-deposition annealing treatment, we deposited some samples at

60 °C and 80 °C and annealed at 120 °C for 90 minutes. Samples were characterized using SEM,

XRD and UV-vis spectroscopy. These samples produced the tetragonal crystal structures with

good intense absorption. And rod-like structures disappear with annealing, fusing together to

make continuous crystalline regions.

We find that the precursor concentration of MACl and PbI2 in DMF and the deposition

temperature are the main factors that affect the formation of highly crystalline pure perovskites

with intense absorption. 0.33 M methylammonium chloride and 0.99 M lead iodide resulted

CH3NH3PbI3-xClx perovskite films with enhanced crystallinity and strong absorption. According

45
to the study, 120 °C is the optimum deposition temperature. Post-film deposition annealing time

is not a factor to increase the crystallinity or the absorption.

Fig. 23. Linear absorption spectra of CH3NH3PbI3-xClx perovskite films at different deposition

temperatures.

46
4. Model to guide the device parameters
Intension of the model is to optimize the nanoparticle concentration to produce the highest

optical absorption and the internal electric field. Higher the volume fraction of the perovskite,

higher the absorbance but, lower the electric field.

The local electric field due to a dipole at a distance r can be written as

 
E dip r  
1
3 p  rˆ rˆ  p 
4 0 r 3

Let’s consider electric field due to nearest neighbor nanoparticles. Assuming a nanoparticle as a

single dipole and considering a cubic volume which has 8 dipoles at the corners,

5 6

2
3

E

8
7

4
1

Fig. 24. Hypothetical cubic arrangement of BTO nanoparticles.


E - field at the center of the cube due to dipole 1, is in following direction.


E - field at the center of the cube due to dipole 6, is in following direction.

47
θ


Therefore r̂ components cancel out leaving only the p component. Therefore, the e-field at the

center of a single cubic volume element is in the poled direction and on the order of eight

nanoparticles electric field.

In terms of polarization, the local electric field due to a nanoparticle at a distance r can be written

as

  R3P
E BTO (r )   pˆ 
3 0 r 3

where the ε0 is the free space permittivity and R is the radius of a nanoparticle.
Then the average electric field magnitude due to eight nanoparticles at a distance of 30 nm from
the center of each nanoparticle is ~ 4 × 1010 Vm-1.
The average electric field magnitude due to eight nanoparticles at a distance of 50 nm from the
center of each particle is ~ 1010 Vm-1.
The average electric field magnitude due to eight nanoparticles at a distance of 75 nm from the
center of each particle is ~ 3 × 109 Vm-1.
R (nm) Electric Field (Vm-1)
30 4 × 1010
50 1010
75 3 × 109

Table 4. Electric field variation with distance.


The volume of the absorber needs to be determined. For that volume fraction of the absorber in a
single cubic volume element can be calculated.

48
2r2 = (d + 2R)2

d + 2R = √2𝑟
d
r
d  2R
r=
2

Fig. 25. Relationship between nanoparticle distance, size of a nanoparticle and the distance from

one nanoparticle to the center of the cube.

Volume contribution to a cube from 8 spherical nanoparticles is equal to the volume of one

nanoparticle.

Let d + 2R = ϒR, R is the radius of a nanoparticle.

V
R 3  4 R 3 4
3 
Volume fraction of the absorber is, nano  3  3  fv
Vcube (R) 3 3

For 50 % volume fraction,


4
3 
fv  3  0.5    2.03098  d  2 R  2.03098R  50.7745nm
3
4
3 
For 60 % volume fraction, fv  3  0.6    2.1878  d  2 R  2.1878R  54.695nm
3
4
3 
For 70 % volume fraction, fv  3  0.7    2.408  d  2 R  2.408R  60.2nm
3
4
3 
For 80 % volume fraction, fv  3  0.8    2.756  d  2 R  2.756R  68.9nm
3
4
3 
For 90 % volume fraction, fv  3  0.9    3.473  d  2 R  3.473R  86.825nm
3

49
4
3 
For 95 % volume fraction, fv  3  0.95    4.376  d  2 R  4.376R  109.39nm
3
fv r (nm)
50 35.903
60 38.675
70 42.568
80 48.72
90 61.394
95 77.35

Table 5. Perovskite absorber volume fraction variation with distance.

E field (Vm-1) Volume Fraction


100 %
1011
90 %
9
10 80 %

70 %

60 %
50%

V.F.

E-field)

25 50 75 distance (nm)

Fig. 26. Electric field versus distance from the center of a nanoparticle and volume fraction of

the absorber versus distance from the center of a nanoparticle.

Therefore a volume fraction of 87% of the perovskite absorber and an E-field of ~ 8 × 109 Vm-1
at a distance of 60 nm from the center of nanoparticles will produce the highest optical
absorption and the internal electric field.

50
Appendix A

A.1. Schematic diagram of the Vibrating-Mesh Nebulizer

Lid Top View

1.221 cm
3.0 cm

2.442 cm O.C. Hole d = 2.2 mm


(screw)

3.5 cm

Side View

0.5 cm

Finish for O-ring seal

51
Cross Sectional View

Black O-ring
Top View

Screws 2X

Electrodes

52
Front View (Electrodes) 3.5 cm
1.01 cm O.C.
1.249 cm

0.306 cm

0.704 cm
3.9 cm

0.402 cm 0.402 cm

0.71 cm
Side View (Electrodes)

0.306 cm

3.9 cm

53
Bottom View

Black O ring

Screws threads

54
References

[1]. William, S.; Hans, J.Q. J. Appl. Phys., 32, 510:1–510:10, 1961.

[2]. M.D. Graef and M. McHenry, Cambridge University Press, 2007.

[3]. D. Weber, Z. Naturforsch., 33b, 1443, 1978.

[4]. C. S. Ponseca, Jr., T. J. Savenije, M. Abdellah, K. Zheng, A. Yartsev, T. Pascher, T.

Harlang, P. Chabera, T. Pullerits, A. Stepanov, J-P. Wolf, V. Sundström, J. Am. Chem.

Soc., 136, 5189−5192, 2014.

[5]. C. Wehrenfennig, G.E. Eperon, M.B. Johnston, H.J. Snaith, L.M. Herz, Adv. Mater,

26, 1584-1589, 2014.

[6]. M. J. Carnie, C. Charbonneau, M. L. Davies, J. Troughton, T. M. Watson, K.

Wojciechowski, H. Snaith, D. A. Worsley, Chem. Commun., 49, 7893, 2013.

[7]. S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. P. Alcocer, T. Leijtens, L.

M. Herz, A. Petrozza, H. J. Snaith , , Science, 342 , 341, 2013.

[8]. A. Kojima, K. Teshima, T. Miyasaka and Y. Shirai, Novel photoelectrochemical cell with

mesoscopic electrodes sensitized by lead-halide compounds (2). Books of Abstracts

(#397), 210th ECS Meeting, Cancun, Mexico, October 29 - November 3, 2006, The

Electrochemical Society, Pennington, New Jersey, 2006.

[9]. H. Zhou, Q. Chen, G. Li, S. Luo, T.-B. Song, H.S. Duan, Z. Hong, J. You, Y. Liu, Y.

Yang, Interface engineering of highly efficient perovskite solar cells, Science, 345, 542,

2014.

[10]. E. Edri, S. Kirmayer, D. Cahen, and G. Hodes, J. Phys. Chem. Lett. 4(6), 897, 2013.

[11]. H. S. Kim, I. Mora-Sero, V. Gonzalez-Pedro, F. Fabregat-Santiago, E. J. Juarez-Perez, N.

G. Park, and J. Bisquert, Nat. Commun. 4, 2242, 2013.

55
[12]. E. Mosconi, A. Amat, M. K. Nazeeruddin, M. Grätzel and F. De Angelis, J. Phys. Chem.

C, 117, 13902, 2013.

[13]. C. C. Stoumpos, C. D. Malliakas and M. G. Kanatzidis, Inorg. Chem. 52, 9019, 2013.

[14]. J.H. Im, J. Chung, S.-J. Kim and N.G. Park, Nanoscale Res. Lett. 7, 353, 2012.

[15]. J. You, Z. Hong, Y. M. Yang, Q. Chen, M. Cai, T. B. Song, C. C. Chen, S. Lu, Y. Liu, H.

Zhou, and Y. Yang, ACS Nano 8(2), 1674, 2014.

[16]. W. A. Laban and L. Etgar, Energy Environ. Sci. 6(11), 3249, 2013.

[17]. G. Xing, N. Mathews, S. Sun, S. S. Lim, Y. M. Lam, M. Gratzel, S. Mhaisalkar, and T.

C. Sum, Science, 342(6156), 344, 2013.

[18]. H.S. Kim, J.W. Lee, N. Yantara, P.P. Boix, S.A. Kulkarni, S. Mhaisalkar, S. Gratzel,

N.G.Park, Nano lett., 13: 2412, 2013.

[19]. M.Z. Liu, M.B. Johnston, H.J. Snaith, Nature, 501:395, 2013.

[20]. Goldschmidt, V. M. Crystal Structure and Chemical Combination. Ber. Dtsch. Chem.

Ges., 60, 1263−1268, 1927.

[21]. C. Li, X. Lu, W. Ding, L. Feng, Y. Gao, Z. Guo, Formability of ABX3 (X = F, Cl, Br, I)

Halide Perovskites. Acta Crystallogr., B64, 702−707, 2008.

[22]. Cohen, B. N. Labarca, C. Davidson, N. Lester, H. A. Mutations In M2 Alter The

Selectivity of the Mouse Nicotinic Acetylcholine Receptor for Organic and Alkali Metal

Cations. J. Gen. Physiol., 100, 373−400, 1992.

[23]. I. Chung, J. H. Song, J. Im, J. Androulakis, C. D. Malliakas, H. Li, A. J. Freeman, J. T.

Kenney, and M. G. Kanatzidis, J. Am. Chem. Soc. 134(20), 8579, 2012.

[24]. T. Ishihara, J. Lumin. 60–61, 269, 1994.

56
[25]. Poglitsch, A. Weber, D. Dynamic Disorder in Methylammonium trihlogenoplumbates(II)

Observed by Millimeter-Wave Spectroscopy. J. Chem. Phys., 87, 6373−6378, 1987.

[26]. J. J. Choi, X. Yang, Z. M. Norman, S. J. L. Billinge, and J. S. Owen, Nano Lett. 14(1),

127, 2014.

[27]. A. Kojima, K. Teshima, Y. Shirai, and T. Miyasaka, J. Am. Chem. Soc. 131(17), 6050,

2009.

[28]. F. Brivio, A. B. Walker, and A. Walsh, APL Mater. 1(4), 042111, 2013.

[29]. T. Baikie, Y. Fang, J. M. Kadro, M. Schreyer, F. Wei, S. G. Mhaisalkar, M. Gr¨atzel, and

T. J. White, J. Mater. Chem. A, 1(18), 5628, 2013.

[30]. T. Baikie, Y. Fang, J. M. Kadro, M. Schreyer, F. Wei, S. G. Mhaisalkar, M. Gr¨atzel, and

T. J. White, Synthesis and Crystal Chemistry of the Hybrid Perovskite (CH3NH3)PbI3 for

Solid-State Sensitised Solar Cell Applications. J. Mater. Chem. A, 1, 6328 – 6341, 2013.

[31]. Wang, S., Mitzi, D. B., Field, C. A., Guloy, A. Synthesis and Characterization of

[NH2Cl):NH2]3MI5(M=Sn,Pb): Stereochemical Activity in Divalent Tin and Lead Halides

Containing Single <110> Perovskite Sheets. J. Am. Chem. Soc., 117, 5297−5302, 1995.

[32]. J. Even, L. Pedesseau, J.M. Jancu, C. Katan, Importance of Spin Orbit Coupling in

Hybrid Organic/Inorganic Perovskites for Photovoltaic Applications. J. Phys. Chem. Lett.

4, 2999−3005, 2013.

[33]. K. Tanaka, T. Takahashi, T. Ban, T. Kondo, K. Uchida, N. Miura, Comparative Study on

the Excitons in Lead-Halide-Based Perovskite-Type Crystals CH3NH3PbBr3

CH3NH3PbI3. Solid State Commun. 127, 619−623, 2003.

57
[34]. G.C. Papavassiliou, G.A. Mousdis, I.B. Koutselas, Some New Organic-Inorganic Hybrid

Semiconductors Based on Metal Halide Units: Structural, Optical and Related Properties.

Adv. Mater. Opt. Electron. 9, 265−271, 1999.

[35]. G.C. Papavassiliou, Three- and LOW-Dimensional Inorganic Semiconductors. Prog.

Solid State Chem. 25, 125−270, 1999.

[36]. G.A. Mousdis, V. Gionis, G.C. Papavassiliou, C.P. Raptopoulou, A. Terzis, Preparation,

Structure and Optical Properties of [CH3SC(=NH2)NH2]3PbI5,

[CH3SC(=NH2)NH2]4Pb2Br8 and [CH3SC(=NH2)NH2]3PbCl5• CH3SC(=NH2)NH2Cl. J.

Mater. Chem., 8, 2259−2262, 1998.

[37]. S. Elleuch, T. Dammak, Y. Abid, A. Mlayah, H. Boughzala, Synthesis, Structural and

Optical Properties of a Novel Bilayered Organic-Inorganic Perovskite C5Pb2I5. J.

Lumin., 130, 531−535, 2010.

[38]. B.A. Gregga, M.C. Hanna, Comparing Organic to Inorganic Photovoltaic Cells: Theory,

Experiment, and Simulation. J. Appl. Phys., 93, 3601−3614, 2003.

[39]. H.S. Kim, I. Mora-Sero, V. Gonzalez-Pedro, F. Fabregat-Santiago, E.J. Juarez-Perez,

N.G. Park, J. Bisquert, Mechanism of Carrier Accumulation in Perovskite Thin-Absorber

Solar Cells. Nat. Commun., 4, 2242, 2013.

[40]. J. M. Frost, K. T. Butler, F. Brivio, C. H. Hendon, M. van Schilfgaarde and A. Walsh,

preprint arXiv:1402.4980v2, 2014.

[41]. J. J. Choi, X. Yang, Z. M. Norman, S. J. L. Billinge, and J. S. Owen, Nano Lett. 14(1),
127 (2014).
[42]. J. Fousek, Proceedings of the Ninth IEEE International Symposium on Applications

of Ferroelectrics, ISAF '94, pp. 1–5, 1994.

58
[43]. M. E. Lines and A. M. Glass, Principles and Applications of Ferroelectrics and Related

Materials, Oxford, UK, Clarendon, 1977.

[44]. J. Valasek, “Piezo-Electric and Allied Phenomena in Rochelle Salt,” Phys. Rev., Vol. 17,

pp. 475-481, 1921.

[45]. P. H. Lindenmeyer and R. Hosemann, “Application of the theory of paracrystals to the

crystal structure analysis of polyacrylonitrile,” J. Appl. Phys., Vol. 34, p. 42, 1963.

[46]. T. Choi, S. Lee, Y. Choi, V. Kiryukhin, & S.W. Cheong, Switchable ferroelectric diode

and photovoltaic effect in BiFeO3, Science, 324, 63-66, 2009.

[47]. S.Y. Yang et al., Above-bandgap voltages from ferroelectric photovoltaic devices. Nature

Nanotechno l.5, 143-147, 2010.

[48]. D. Cao et al., High-efficiency ferroelectric-film solar cells with an n-type Cu2O cathode

buffer layer. Nano Lett. 12, 2803-2809, 2012.

[49]. M. Alexe, D. Hesse, Tip-enhanced photovoltaic effects in bismuthferrite, Nature

Commun.2, 256(2011).

[50]. M. Qin, K. Yao, Y.C. Liang, High efficiency photovoltaics in nano scaled ferroelectric

thin films. Appl. Phys. Lett. 93, 122904, 2008.

[51]. W.S. Choi et al., Wide bandgap tunability in complex transition metal oxides by site-

specific substitution. Nature Commun. 3,689, 2012.

[52]. J. Kreisel, M. Alexe, P.A. Thomas, A photo ferroelectric material is more than the sum of

its parts. Nature Mater.11, 260, 2012.

[53]. S. Liu, F. Zheng, N.Z. Koocher, H. Takenaka, F. Wang, and A.M. Rappe, J. Phys. Chem.

Lett., 6, 693-699, 2015.

59
[54]. Y. J. Park, I.S. Bae, S. J. Kang, J. Chang and C. Park, IEEE Trans. Dielectr. Electr.

Insul., 17, 1135-1163, 2010.

[55]. I. B. Koutselas, L. Ducasse and G. C. Papavassiliou, J. Phys.: Condens. Matter, 8,

1217 - 1227, 1996.

[56]. J. M. Gregg, Phys. Status Solidi A, 206, 577–587, 2009.

[57]. N. A. Spaldin, Science, 304, 1606 –1607, 2004.

[58]. P. Yung, L. Won-Jae and K. Ho-Gi, J. Phys.: Condens. Matter., 9, 9445, 1997.

[59]. C. H. Ahn, K. M. Rabe and J.-M. Triscone, Science, 303,488–491, 2004.

[60]. D. D. Fong, G. B. Stephenson, S. K. Streiffer, J. A. Eastman,O. Auciello, P. H. Fuoss

and C. Thompson, Science, 304, 1650–1653, 2004.

[61]. I. I. Naumov, L. Bellaiche and H. Fu, Nature, 432, 737–740, 2004.

[62]. G. Geneste, E. Bousquet, J. Junquera and P. Ghosez, Appl. Phys. Lett., 88, 112906,

2006.

[63]. J. Hong and D. Fang, Appl. Phys. Lett., 92, 012906, 2008.

[64]. B. Vilquin, B. Gautier, A. Brugere and J. S. Moulet, AIP Conf. Proc., 1173, 129–134,

2009.

[65]. G. Cao, Nanostructures and Nanomaterials: Synthesis, Properties, and

Applications, Imperial College Press, London, 1 edn, 2004.

[66]. B. Hilczer J. Kulek, M. Polomska M.D. Glinchuk A.V. Ragulya, A. Pietraszko A,

Ferroelectr, 316:31, 2005.

[67]. D. Lijie, X. Chuanxi, C.H. Juan, N. Cewen, J Wuhan Univ Technol Mater Sci, Ed 19:9,

2004.

[68]. S.J. Gallagher, B. Norton, P.C. Eames, Sol Energy, 81:813, 2007.

60
[69]. Z.M. Dang, H.Y. Wang, Y.H. Zhang, J.Q. JQ, Macromol Rapid Commun 26:1185, 2005.

[70]. Q. Hongying, X. Chuanxi, D. Lijie, Z. Guanghui, J Wuhan, Univ Technol Mater Sci Ed

21:133, 2006.

[71]. L. Yan, K. Wang, L. Ye, J. Mater Sci. Lett. 22:1713, 2003.

[72]. Y. P. Mao, S. Y. Mao, Z.-G. Ye, Z. X. Xie, and L. S. Zheng, J. Appl. Phys. 108, 014102,

2010.

[73]. D. Ginley, M.A. Green, R. Collins, Solar energy conversion toward 1 terawatt. Mater.

Res. Soc. Bull. 33, 355–364, 2008.

[74]. J. Seidel, D. Fu, S. Y. Yang, E. Alarcon-Llado, J. Wu, R. Ramesh, and J. W. Ager III,

Phys. Rev. Lett. 107, 126805, 2011.

[75]. Y. M. Sheu, S. A. Trugman, Y.S. Park, S. Lee, H. T. Yi, S.-W. Cheong, Q. X. Jia, A. J.

Taylor, and R. P. Prasankumar, Appl. Phys. Lett. 100, 242904, 2012.

[76]. R. Von Baltz, Phys. Status Solidi B, 89, 419 - 429, 1979.

[77]. A. G. Chynoweth, Phys. Rev., 102, 705 - 714, 1956.

[78]. D. Ginley, M.A. Green, R. Collins, Solar energy conversion toward 1 terawatt. Mater.

Res. Soc. Bull. 33, 355–364, 2008.

[79]. I. Gur, N.A. Fromer, M.L. Geier, A.P. Alivisatos, Air-stable all-inorganic nanocrystal

solar cells processed from solution. Science, 310, 462 - 465, 2005.

[80]. B. O'Regan, M.A. Grätzel, A low-cost, high-efficiency solar cell based on dye-sensitized

colloidal TiO2 films. Nature, 353, 737–740, 1991.

[81]. K. Yao, B. K. Gan, M. Chen, and S. Shannigrahi, Appl. Phys. Lett. 87, 212906, 2005.

[82]. Y.S. Yang, S.L. Lee, S. Yi, B.G. Chae, S.H. Lee, H.J. Joo, M.S. Jang, Appl. Phys. Lett.

76:774, 2000.

61
[83]. F.G. Zheng, J. Xu, L. Fang, M.R. Shen, X.L. Wu, Appl. Phys. Lett. 93:172101, 2008.

[84]. L. Pintilie, I. Vrejoiu, G.L. Rhun, M. Alexe, J. Appl. Phys. 101:064109, 2007.

[85]. B. Chen, M. Li, Y.W. Liu, Z.H. Zuo, F. Zhuge, Q.F. Zhan, R.W. Li, Nanotechnology

22:195201, 2011.

[86]. Y.Y. Zang, D. Xie, X. Wu, Y. Chen, Y.X. Lin, M.H. Li, H. Tian, X. Li, Z. Li, H.W. Zhu,

T.L. Ren, D. Plant, Appl. Phys. Lett. 99:132904, 2011.

[87]. J.R. Teague, R. Gerson, and W.J. James, Solid State Commun., 8, 1073, 1970.

[88]. J. Wang, J.B. Neaton, H. Zheng, V. Nagarajan, S.B. Ogale, B. Liu, D. Viehland, V.

Vaithyanathan, D.G. Schlom, U.V. Waghmare, N.A. Spaldin, K.M. Rabe, M. Wuttig, and

R. Ramesh, Science, 299,1719-1722, 2003.

[89]. K.Y. Yun, D. Ricinschi, T. Kanashima, M. Noda, and M. Okuyama, Jap. J. Appl. Phys.,

43, L647, 2004.

[90]. S.K. Singh and H. Ishiwara, Jap. J. Appl. Phys., 44, L734, 2005.

[91]. G. Catalan, J.F. Scott. Physics and applications of bismuth ferrite. Adv. Mater, 21(24):

2463 - 2485, 2009.

[92]. T.L. Qu, Y.G. Zhao, D. Xie, J.P. Shi, Q.P. Chen, T.L. Ren, Appl. Phys. Lett. 98:173507,

2011.

[93]. T. Choi, S. Lee, Y.J. Choi, V. Kiryukhin, S.W. Cheong, Science. 324:63, 2009.

[94]. W. Ji, K. Yao, Y.C. Liang, Adv. Mater 22:1763, 2010.

[95]. C.S. Tu, C.M. Hung, V.H. Schmidt, R.R. Chien, M.D. Jiang, J. Anthoninappen, J. Phys.:

Condens. Matter. 24:495902, 2012.

[96]. C.M. Hung, M.D. Jiang, J. Anthoninappen, C.S. Tu, J. Appl. Phys., 113:17D905, 2013.

[97]. R. Guo, L. You, L. Chen, D. Wu, J.L. Wang, Appl. Phys. Lett. 99:122902, 2011.

62
[98]. H.T. Huang, Nat. Photonics. 4:134, 2010.

[99]. H.W. Chang, F.T. Yuan, Y.C. Yu, P.C. Chen, C.R. Wang, C.S. Tu, S.U. Jen, J. Alloys.

Compd. 574:402, 2013.

[100]. Y.Y. Zang, D. Xie, Y. Chen, X. Wu, T.L. Ren, J.Q. Wei, H.W. Zhu, D. Plant, Nanoscale,

4:2926, 2012.

[101]. W. Dong, Y.P. Guo, B. Guo, H.Y. Liu, H. Li, H.Z. Liu, Mater. Lett., 91:359, 2013.

[102]. P. Poosanaas, K. Tonooka and K. Uchino, Mechatronics, 10, 467-487, 2000.

[103]. P. Brody and B. Rod, Integr. Ferroelectr., 3, 245-257, 1993.

[104]. A. Matsumura, Y. Kamaike, T. Horiuchi, M. Shimizu, T. Shiosaki and K. Matsushige,

Jpn. J. Appl. Phys., Part 1, 34, 5258-5262, 1995.

[105]. R. Katiyar, A. Kumar, G. Morell, J. Scott and R. Katiyar, Appl. Phys. Lett., 99, 092906,

2011.

[106]. Y. Yao, B. Zhang, L. Chen, Y. Yang, Z. Wang, H. N. Alshareef and X. X. Zhang, J.

Phys. D: Appl. Phys., 46, 055304, 2013.

[107]. M. Hirasawa, T. Ishihara and T. Goto, J. Phys. Soc. Jpn.,63, 3870 – 3879, 1994.
[108]. X. Deng, X. Wang, H. Wen, A. Kang, Z. Gui, and L. Li, J. Am. Ceram. Soc. 89, 1059

2006.

[109]. R. Kretschmer and K. Binder, Phys. Rev. B: Condens. Matter Mater. Phys., 20, 1065 -

1076, 1979.

[110]. W. Lee, H. Han, A. Lotnyk, M. A. Schubert, S. Senz, M. Alexe, D. Hesse, S. Baik and U.

Gosele, Nat. Nanotechnol., 3, 402 - 407, 2008.

[111]. M. Alexe, C. Harnagea, D. Hesse and U. Gosele, Appl. Phys. Lett., 75, 1793 -1795, 1999.

[112]. M. Alexe, C. Harnagea, A. Visinoiu, A. Pignolet, D. Hesse and U. Gosele, Scr. Mater.,

44, 1175 - 1179, 2001.

63
[113]. M. Dawber, P. Chandra, P. B. Littlewood and J. F. Scott, J.Phys.: Condens. Matter, 15,

L393, 2003.

[114]. Y. Li, Z. Liao, F. Fang, X. Wang, L. Li, and J. Zhu, Appl. Phys. Lett. 105, 182901, 2014.

[115]. J. Zhu, W. Han, H. Zhang, Z. Yuan, X. Wang, L. Li, and C. Jin, J. Appl. Phys. 112,

064110, 2012.

[116]. M. Yashima, T. Hoshina, D. Ishimura, S. Kobayashi, W. Nakamura, T.

Tsurumi,S. Wada, J. Appl. Phys. 98, 014313, 2005.

[117]. C.H. Kim, K.J. Park, Y.J. Yoon, D.S. Sinn, Y.T. Kim, K.H. Hur, J. Eur.

Ceram. Soc.28, 2589, 2008.

[118]. T. M. Shaw, S. Trolier-McKinstry and P. C. McIntyre, Annu. Rev. Mater. Sci., 30, 263 -

298, 2000.

[119]. C.J. Xiao, Z.H. Chi, W.W. Zhang, F.Y. Li, S.M. Feng, C.Q. Jin, X.H. Wang,

X.Y. Deng, L.T. Li, J. Phys. Chem. Solids, 68, 311, 2007.

[120]. S. Wada, T. Hoshina, H. Yasuno, S.M. Nam, H. Kakemoto, T. Tsurumi,

M. Yashima, J. Korean Phys. Soc. 46, 303, 2005.

[122]. S. Wada, M. Narahara, T. Hoshina, H. Kakemoto, T. Tsurumi, J. Mater. Sci.

38, 2655, 2003.

[123]. S. Wada, H. Yasuno, T. Hoshina, S.-M. Nam, H. Kakemoto, and T. Tsurumi, Jpn. J.

Appl. Phys., Part 1 42, 6188, 2003.

[124]. X.H. Wang, H. Wen, R.Z. Chen, H. Zhou, L.T. Li, Size effect of barium

titanatebased ceramics sintered in reducing atmospheres, in: Proceedings of

the 50th Anniversary Conference on IEEE International Ultrasonics,

Ferroelectrics, and Frequency Control Joint, 2004.

64
[125]. Y. G. Wang, W. L. Zhong and P. L. Zhang, Solid State Commun., 92, 519 - 523, 1994.

[126]. T. Hoshina, S. Wada, Y. Kuroiwa, T. Tsurumi, Appl. Phys. Lett. 93,192914, 2008.
[127]. M. J. Polking, M.-G. Han, A. Yourdkhani, V. Petkov, C. F. Kisielowski, V. V. Volkov,

Y. Zhu, G. Caruntu, A. P. Alivisatos and R. Ramesh, Nat. Mater., 11, 700 -709, 2012.

[128]. C. Pithan, Y. Shiratori, R. Waserz, J. Dornseiffer, F.H. Haegel, J. Am. Ceram.

Soc.89, 2908, 2006.

[129]. Z. Zhao, V. Buscaglia, M. Viviani, M.T. Buscaglia, L. Mitoseriu, A. Testino,

M. Nygren, M. Johnsson, P. Nanni, Phys. Rev. B 70,024107, 2004.

[130]. H. Zhang, X. Wang, Z. Tian, C. Zhong, Y. Zhang, C. Sun, and L. Li, J. Am.

Ceram. Soc. 94, 3220, 2011.

[131]. H. S. Kim, C. R. Lee, J. H. Im, K. B. Lee, T. Moehl, A. Marchioro, S. J. Moon,

R. Humphry-Baker, J. H. Yum, J. E. Moser, M. Gratzel and N. G. Park, Sci. Rep., 2,

591, 2012.

[132]. J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and

M. Grätzel, Nature, 499, 316 - 319, 2013.

[133]. M. Liu, M.B. Johnston, H.J. Snaith, , Nature, 501, 395-398, 2013.

[134]. Q. Chen, H. Zhou, Z. Hong, S. Luo, H.S. Duan, H.H. Wang, Y. Liu, G. Li, Y. Yang, J.

Am. Chem. Soc., 136, 622-625, 2014.

[135]. M. A. Green, Solar cells: operating principles, technology, and system applications,

Englewood Cliffs, NJ, Prentice-Hall, Inc., 1982.

[136]. S. Colella, E. Mosconi, P. Fedeli, A. Listorti, F. Gazza, F. Orlandi, P. Ferro, T. Besagni,

A. Rizzo, G. Calestani, G. Gigli, F. De Angelis, R. Mosca, Chem. Mater., 25, 4613 -

4618, 2013.

65
[137]. Y. Zhao, K. Zhu, CH3NH3Cl - Assisted One-Step Solution Growth of CH3NH3PbI3

Structure, Charge - Carrier Dynamics, and photovoltaic Properties of Perovskite Solar

Cells. J. Phys. Chem. C, 118, 9412-9418, 2014.

[138]. S.R. Basu, L.W. Martin, Y.H. Chu, M. Gajek, R. Ramesh, R.C. Rai, X. Xu, J.L.

Musfeldt, Appl. Phys. Lett. 92:091905, 2008.

[139]. D. Cahen, G. Hodes, M. Gratzel, J.F. Guillemoles, I. Riess, J. Phys. Chem. B, 104:2053,

2000.

[140]. M. Turrion, J. Bisquert, P. Salvador, J. Phys. Chem. B 107:9397, 2003.

[141]. X. Wei, T,F. Xie, D. Xu, Q.D. Zhao, S. Pang, D.J. Wang, Nanotechnology 19:275707,

2008.

[142]. S.J. Clark, J. Robertsona, Appl. Phys. Lett. 90:132903, 2007.

[143]. T. Ishida, H. Kobayashi, Y. Naano, J. Appl. Phys. 73:4344, 1993.

[144]. C. Wang, K.J. Jin, Z.T. Xu, L. Wang, C. Ge, H.B. Lu, H.Z. Guo, M. He, G.Z. Yang,

Appl. Phys. Lett. 98:192901, 2011.

[145]. G.L. Yuan, J. Wang, Appl. Phys. Lett. 95:252904, 2009.

[146]. Y. Zhou, B. Yu, X. Zhu, X. Tan, L. Qian, L. Liu, J. Yu, S. Yuan, J. Sol-Gel. Sci.

Technol., 72:74 - 79, 2014.

[147]. M. Qin, K. Yao, Y. C. Liang and S. Shannigrahi, J. Appl. Phys., 101, 014104 - 014108,

2007.

[148].L. Huang, Z.Y. Chen, J.D. Wilson, S. Banerjee, R.D. Robinson, I.P. Herman,

R. Laibowitz, S.O. Brien, J. Appl. Phys. 100, 034316, 2006.

66
Publications and Presentations

Publications
1. Chaminda Hettiarachchi, Nicholas Valdes, Pritish Mukherjee, Sarath Witanachchi, “A
novel single-step growth process for the deposition of CH3NH3PbI3-xClx perovskite films
from CH3NH3Cl and PbI2 precursors”, The journal of Material Research, has been
submitted.

2. Julia E. Medvedeva and Chaminda L. Hettiarachchi, “Tuning the properties of complex


transparent conducting oxides: Role of crystal symmetry, chemical composition, and
carrier generation”, Physical Review B, 81, pg.125116, (2010).

Oral Presentations
1. Chaminda Hettiarachchi, Nicholas Valdes, Pritish Mukherjee, Sarath Witanachchi, “A
novel growth process for the deposition of CH3NH3PbI3-xClx perovskite films”, 2014
Energy Materials Nanotechnology (EMN) Fall Meeting, DoubleTree by Hilton Hotel,
Orlando, Florida, USA, November 22 to 25, 2014.

2. Chaminda Hettiarachchi, Robert Hyde, Samantha Wijewardena, Pritish Mukherjee and


Sarath Witanachchi, “ Structural and Optical Properties of PbSe quantum dots treated
with 3-mercaptopropionic acid”, Florida Annual Meeting & Exposition, Palm Harbor
Florida, May 2013.

3. Chaminda Hettiarachchi, Domingo Feliciano, Devjyoti Mukherjee, Pritish Mukherjee and


Sarath Witanachchi, “Improvement of carrier transport in PbSe quantum dot-embedded
polymeric solar cells fabricated by a Laser Assisted Spray process”, American Vacuum
Society 59th International Symposium & Exhibition, Tampa, FL, October 28-November 2,
2012.

Poster Presentations

1. Chaminda Hettiarachchi, Nicholas Valdes, Pritish Mukherjee, Sarath Witanachchi, “A


novel growth process for the deposition of CH3NH3PbI3-xClx perovskite films”, Materials
Research Society Spring Meeting, San Francisco CA, April 2015.

2. Chaminda Hettiarachchi, Marek Merlak, Pritish Mukherjee and Sarath Witanachchi,


“Investigation of Multiple Exciton Generation - Dissociation in PbSe Quantum Dots
Embedded in a PbTe Matrix”, Materials Research Society Spring Meeting, San Francisco
CA, April 2014.

67
3. Chaminda Hettiarachchi, Domingo Feliciano, Robert Hyde, Pritish Mukherjee and
Sarath Witanachchi, “Investigation of carrier transport in surfactant-free PbSe quantum
dot embedded bulk heterojuncion polymeric solar cells fabricated by a Laser Assisted
Spray process”, Materials Research Society fall Meeting, Boston MA, October 2012.

4. Chaminda Hettiarachchi, Dino Ferizovic, Devjyoti Mukherjee, Robert Hyde, Pritish


Mukherjee and Sarath Witanachchi, “Structural and Optical Properties of Surfactant-free
Coatings of PbSe Quantum Dots deposited by a Laser Assisted Spray Process”,
International Conference on Surfaces, Coatings and Nanostructured Materials
(NANOSMAT-USA), Tampa FL, 27-30 March 2012.

5. , Chaminda Hettiarachchi, Domingo Feliciano, Robert Hyde, Pritish Mukherjee and


Sarath Witanachchi,” Investigation of multiexciton generation-dissociation in surfactant-
free PbSe quantum dots”, Materials Research Society Spring Meeting, San Francisco
CA, April 2013.

Other Research Accomplishments

 Guided REU (Research Experience for Undergraduates) student Constant Owens to win
the second place of the 2012 summer REU program, titled “Growth and characterization
of quantum dot-carbon nanotube composite structures for excitonic solar cells”.

 Guiding undergraduate student Nicholas Valdes towards honors thesis on “A noval


approach for fabrication and characterization of organolead halide perovskite solar cells”.

68

View publication stats

You might also like