You are on page 1of 82

Sequential Deposition Method for Planar Perovskite Solar Cells

Graduation report of
Junke Wang

Supervisor
Ir. B.J. Bruijnaers

Supervising Professor
Prof. Dr. Ir. R.A.J. Janssen

Advising Committee
Dr. Ir. M.M. Wienk
Dr. M.G. Debije

Eindhoven, April 2016


Molecular Materials and Nanosystems
Laboratory of Macromolecular and Organic Chemistry
Department of Chemical Engineering and Chemistry
Eindhoven University of Technology

Abstract
Organometal halide perovskite solar cells are very appealing alternative photovoltaic technologies
for the future solar energy market. In possessing the advantages of inexpensive feedstocks,
solution processability, and large scalability, perovskite solar cells have captured considerable
attention in the organic photovoltaic (OPV) research field, with an unprecedentedly rapid growth
in power conversion efficiencies from 3.8% to 21% in six years. In this study, research has been
focused on the fabrication of simplified, low temperature processed (T < 150 C) planarstructured p-i-n perovskite devices by employing a feasible sequential deposition technique, and
to gain a profound understanding of the formation processes of perovskite films.
In Chapter 2, the solution-processed sequential deposition approach is investigated, which
involves the deposition of a lead halide precursor film in the first step, followed by a chemical
reaction with the organic ammonium halide in a solution phase, to achieve the conversion from
the inorganic precursor into the hybrid perovskite crystals. Here, compact PbI2, porous PbI2 and
mixed lead halide (PbI2 and PbCl2) precursor films are fabricated, and the impacts of these
precursor compositions and morphologies on the conversion into perovskite have been
systematically investigated. It is found that all three precursors suffer from the limited conversion
of the lead source, which is induced by the hampered infiltration of the organic solution phase.
The incomplete conversion leads to a deficient thickness of the light absorber layer and impairs
the device performance. After optimizing the second-step CH3NH3I deposition and postannealing treatment, under the standard test condition, a device made from the compact PbI2
gives a power conversion efficiency (PCE) of 7.52%, while the PCEs derived from the porous
PbI2 and mixed lead halide films are 8.95% and 8.04%, respectively.
To circumvent the side effects arose from the solution phase, the vapor-assisted deposition
method is studied in Chapter 3, in the hope of achieving a milder reaction condition and better
control over the reaction process. A thermal gradient sublimer is employed for generating the
organic vapor phase, followed by an in-situ reaction between the dilute vapor and the solid predeposited inorganic precursor film in the deposition zone. However, due to the lack of precise
temperature control of the lead halide substrates, it is found that the physical deposition rate is
over-rapid while the chemical reaction is too slow, which severely hinders the complete
conversion. Under the standard test condition, the device fabricated from a compact and a porous
PbI2 displays PCEs of 7.65% and 7.42%, respectively.
To gain a deeper understanding of the intrinsic processes such as recombination mechanisms and
charge transport at selective contacts when operating the perovskite solar cells, electrochemical
impedance spectroscopy (EIS) is carried out in Chapter 4. High-performing perovskite solar cells
are characterized regarding their active layer thicknesses and hole-transporting layer thicknesses.
It is found that a too thick perovskite film correlates to a large density of ionic defects, which
contribute to both the build-up of slow-responded charges within the perovskite film at the lowfrequency, and the increased trap-assisted recombination processes, consistent with its poor PV
performance. Also, it is the quality of the perovskite layer that regulates the recombination
dynamics of the solar cell, while it is found that the change in the PEDOT:PSS layer thickness
does not contribute to different recombination regimes.

iii

Table of contents
Abstract ......................................................................................................................................... iii
Table of contents ........................................................................................................................... iv
Chapter 1 Introduction ................................................................................................................. 1
Abstract ....................................................................................................................................... 1
1.1 Energy concerns and sustainable development ..................................................................... 2
1.2 Classification of solar cells .................................................................................................... 3
1.3 Perovskite solar cells ............................................................................................................. 4
1.3.1 Perovskite: crystal structures and materials properties ................................................... 4
1.3.2 Evolution of perovskite device structure ........................................................................ 6
1.3.3 Characterization of perovskite solar cells ....................................................................... 8
1.3.4 Deposition techniques for planar PSCs and current issues ........................................... 10
1.4 Outline ................................................................................................................................. 13
1.5 References ........................................................................................................................... 14
Chapter 2 Solution-processed sequential deposition for planar perovskite solar cells ......... 16
Abstract ..................................................................................................................................... 16
2.1 Introduction ......................................................................................................................... 17
2.2 Crystallized PbI2 precursor film .......................................................................................... 17
2.2.1 Effect of CH3NH3I concentration on the device performance ...................................... 18
2.2.2 Effect of loading time on the PbI2 conversion .............................................................. 21
2.2.3 Effect of annealing time on device performance .......................................................... 23
2.2.4 Optimized device from c-PbI2 film .............................................................................. 24
2.3 Porous nano-crystallized PbI2 film ...................................................................................... 25
2.3.1 Morphology of p-PbI2 films ......................................................................................... 25
2.3.2 Reaction rate between p-PbI2 films with CH3NH3I ...................................................... 26
2.3.3 Effect of porosity on the reaction rate .......................................................................... 28
2.3.4 Photovoltaic performance of devices made of PbI2120 L TBP precursor ................ 29
2.4 Mixed halide perovskite CH3NH3PbI3-xClx film .................................................................. 31
2.4.1 Effect of Cl content on the formation of lead halide precursor films .......................... 31
2.4.2 Conversion of mixed halide precursor into perovskite ................................................. 33
2.4.3 Photovoltaic performance of mixed halide perovskite devices .................................... 34
2.5 Conclusions ......................................................................................................................... 37
iv

2.6 Experiments ......................................................................................................................... 38


2.7 References ........................................................................................................................... 41
Chapter 3 Vapor-assisted solution process for planar perovskite solar cells......................... 43
Abstract ..................................................................................................................................... 43
3.1 Introduction ......................................................................................................................... 44
3.2 VASP using a thermal gradient sublimer ............................................................................ 44
3.2.1 Vapor deposition on c-PbI2 (compact) substrates ......................................................... 46
3.2.2 Vapor deposition on p-PbI2 (porous) substrates ........................................................... 49
3.3 VASP with home-build equipment ..................................................................................... 53
3.4 Conclusions ......................................................................................................................... 53
3.5 Experiments ......................................................................................................................... 54
3.6 References ........................................................................................................................... 55
Chapter 4 Characterization of perovskite solar cells by electrochemical impedance
spectroscopy ................................................................................................................................. 56
Abstract ..................................................................................................................................... 56
4.1 Introduction ......................................................................................................................... 57
4.2 Perovskite layer thickness ................................................................................................... 59
4.3 PEDOT:PSS layer thickness................................................................................................ 64
4.4 Conclusions ......................................................................................................................... 66
4.5 Experiments ......................................................................................................................... 67
4.6 References ........................................................................................................................... 68
Summary ...................................................................................................................................... 69
Appendix 1 ................................................................................................................................... 71
Appendix 2 ................................................................................................................................... 72
Appendix 3 ................................................................................................................................... 74
Acknowledgements ...................................................................................................................... 76

Chapter 1 Introduction

Abstract
A new generation perovskite solar cells has a potential to get a firm foothold in the future energy
market, by combining promising lowered fabricating costs and high energy conversion
efficiencies. In this chapter, an elaborate introduction about perovskite solar cells is presented.
The formation process of the inorganic-organic hybrid crystal structure and the material
properties are briefly introduced. The evolution of perovskite device structures and the charge
transfer mechanisms are addressed while the challenges in fabricating efficient solar cells and
prospects of various deposition techniques are deliberated. In the end, the objectives of this work
are summarized.

Chapter 1

1.1 Energy concerns and sustainable development


Energy is the currency of science and technology and the foundation for the survival of humanity.
With the rapid growth of population as well as economic development, the global demand for
energy is expected to increase by 58% between 2012 and 2050, from 19 terawatts (TW) to 30
TW.1 This is in contrast to the depletion of the world energy reserves of the non-renewable fossil
resources. Also, ever since the industrial revolution, the massive use of fossil fuels (e.g. coal, oil,
and gas) has been widely blamed for the serious environmental concerns, such as global warming,
stratospheric ozone depletion, and acid precipitation. In the foreseeable future, the only solution
to achieve sustainable development is to make much more use of renewable energy sources and
technologies, while considering the economic profit, reliability, engineering practicality and
public recognition.2 By 2013, renewable energies such as solar, wind, geothermal and
hydropower together with (traditional) biomass, contributed 19.1% to the global final energy
consumption, as can be seen in Fig. 1.1.3 In fact, all of these renewable energy supplies on earth
are derived from the sun, which provides us with a continuous stream of energy that can be
utilized without polluting the environment. Using the photovoltaic effect, one efficient way to
store the solar energy is by transferring it into electric energy. Upon the illumination of sunlight,
photon-generated electrons and holes in materials are swept across the built-in electric field
which creates electrical current and voltage. Such devices are called solar cells.

Figure 1.13 Renewable energy share of the global final energy consumption, 2013.

Figure 1.23 Global installed PV capacity, by the year of 2014.

Introduction

With decades of R&D and vigorous market growth, solar cells are now on the way to becoming
the major energy source in powering the world. The installed global total PV capacity has already
exceeded 177 GW (shown in Fig. 1.2), which produces over 200 TWh of electricity each year and
equally covers the energy supply for over 45 million European households.3, 4 Moreover, it is
estimated that by the end of 2030, the solar PV energy will be able to provide more than 30% of
the electricity demand, among the largest of the total renewable power generation.

1.2 Classification of solar cells


Generally, the selection of light absorber materials in solar cells is based on several factors: (1)
An optimum optical-transition bandgap, to achieve the balance between the absorption of more
above-bandgap photons and less thermalization loss; (2) high quantum efficiency, to convert as
much as absorbed photons to electrons, and generate less non-radiative recombination loss; (3)
cheap and abundant raw materials; (4) stable output efficiency.

Figure 1.35 Best research cell efficiencies, 2016.

Nowadays, mainstream products in the PV market are based on silicon solar cells, including
single crystalline, multi-crystalline and amorphous silicon solar cells. After decades of
investigations, in 2009, single crystalline silicon solar cells attained to a 25% efficiency milestone,
while it was less than 1% in 1941.6 In the meantime, multi-crystalline silicon cells also surpassed
an efficiency of 20%. The crystalline silicon cell has an indirect bandgap, which needs a thick
layer to generate sufficient light absorption. However, the fabrication of a high-purity, thick
silicon film is elaborate and costly, limiting its potential for large-scale production. Lower-cost
amorphous silicon cells have achieved efficiencies around 10%, but the instability and relatively
low efficiencies limit its broad applications. III-V thin film solar cells such as Gallium arsenide
(GaAs), Cadmium telluride (CdTe), Copper indium gallium (di)selenide (CIGS) have also
achieved efficiencies comparable to crystalline silicon cell, however, the toxicity and scarcity of
some elements restrict their industrialization production. Besides inorganic materials, research
has also been conducted for the low-cost organic photovoltaic (OPV) devices.7 Due to its
inexpensive, solution processable and flexible characters, organic solar cells are a very promising
alternative for solar energy conversion. Recently, perovskite solar cells, a new generation of
3

Chapter 1

inorganic-organic hybrid solar cells has attracted considerable attention, by showing the most
rapid increase in the power conversion efficiencies among all PV technologies in history, as can
be seen in Fig. 1.3.

1.3 Perovskite solar cells


Perovskite, named after a Russian mineralogist Lev Perovski, represents any material that has the
same type of crystal structure as calcium titanate (CaTiO3). In nature, there are hundreds of
materials following the perovskite structure and possessing properties such as ferroelectric,
dielectric, piezoelectric, semiconducting and superconducting. Those materials can be either
fabricated via the solid-state sintering at high temperatures or processed at a low temperature in
solution phase, which is promising for printable semiconductors.8 The organometal halide
CH3NH3MX3 (M = Pb or Sn, X = Cl, Br or I) perovskite structure was first reported in 1978,
considering its potential applications for superconductors,9 however, the photovoltaic function of
such inorganic-organic hybrid compounds had never been investigated. In 2009, for the first time,
nano-crystalline particles of CH3NH3PbX3 (X = Br, I) was used as light-sensitizers in dyesensitized solar cells, in achieving a power conversion efficiency (PCE) of 3.8%.10 Ever since
then, hybrid perovskite solar cells have been enthusiastically investigated, creating an
unprecedentedly rapid growth in PCEs that have been reported before,6 from the original 3.8% to
more than 21%, within only six years.

1.3.1 Perovskite: crystal structures and materials properties

Figure 1.4 Representative cubic perovskite crystal unit cell, where A is usually larger than B. 11

Fig. 1.4 shows the typical crystal lattice of the perovskite structure, with a general formula of
ABX3.11 In the unit cell, B-ions are surrounded by a 6-fold of X ions to form the corner-sharing
octahedron network, while an A-ion is in 12-fold cuboctahedral coordination, located in the
center of eight adjacent BX6 octahedra. In hybrid perovskite solar cells, A represents organic
ammonium cation (CH3NH3+, NH=CHNH3+), B represent IV group cations (Pb2+, Sn2+), and X
are halide anions (I-, Br-, Cl- or a mixture of them). The crystal structure of a perovskite material
shows a temperature dependence. At above 54 C, the most commonly studied methylammonium
lead tri-iodide, CH3NH3PbI3, possesses an ideal cubic phase. In between 54 C and -112 C, the
material displays a stabilized tetragonal crystal structure by distorting the PbI2 octahedra along
the c-axis, below which the perovskite crystals stay in an orthorhombic phase.12, 13 The stability of
the tetragonal phase can be predicted by Goldschmidts tolerance factor t (see Equation 1.1).

Introduction

When t lies in the range of 0.81-1.11, the perovskite phase is stable in the cubic phase, while a
less symmetric tetragonal structure appears when 0.81 < t < 0.89 or 1.0 < t < 1.11.12

t ( rA rX ) / (21/2 [rB rX ])

(1.1)

At room temperature, rA = 0.18 nm, rB = 0.119 nm and rX = 0.220 nm (representing the radius of
CH3NH3+, Pb2+ and Cl- ions, respectively),11 the tolerance factor t is therefore 0.83, and the
CH3NH3PbI3 crystal display a tetragonal phase.
To fabricate perovskite film, a commonly used method is the one-step solution deposition, in
which a stoichiometric mixture of PbI2 and CH3NH3I are dissolved in a polar solvent N,Ndimethylformamide (DMF), followed by a spin-coating and annealing process to form
CH3NH3PbI3 crystallites. A typical crystallization process is illustrated in Fig. 1.5. 13 Before
dissolving in DMF, the original PbI2 possesses a layered edge-sharing octahedron structure, with
the interlayer spacing of 6.98 . Upon dissolution and spin coating, CH3NH3+ ions together with
DMF molecules intercalate into the PbI2 network, which expands the interlayer spacing to 11.18
. Due to a screening effect and the competitive bonding between DMF and Pb2+ (via Pb-O
bond), the precipitation of perovskite crystals is suppressed before thermal annealing and
therefore, realizing a better control over the crystallization process. After thermal annealing,
residual DMF molecules evaporates, and CH3NH3I will then react with the PbI2, resulting in the
transformation of [PbI6]4- from the edge-sharing octahedron structures into the corner-sharing
octahedral structures in perovskite crystals, as is shown in Fig. 1.5(c).

Figure 1.513 The crystallization process during the one-step spin-coating deposition. (a) The PbI2 crystallites before
dissolution in DMF solvent. (b) After dissolution and spin-coating, the layer spacing of Pb-I network is expanded
which facilitates the intercalation of CH3NH3+ ions. (c) After thermal annealing, CH3NH3+ reacts with PbI2 in the
precipitation of perovskite crystals.

It is clear that the simplicity of this fabrication method as well as the vast availability of the
chemical feedstocks, appeal to many researchers to devote to perovskite PV devices. Through
either solution or vacuum processes at low temperatures, an ultrathin perovskite film (300-400
nm)14 sandwiched between two organic layers can be directly deposited onto a flexible polymer
substrate, resulting in a PCE higher than 10% and maintain good performance after a bending
test.15, 16 This preliminary success demonstrates the possibility of mass-production of efficient
perovskite solar devices by roll-to-toll printing technique. With only several hundreds of
nanometers (ca. 300 nm) active layers, efficiencies such as 15.4% can be easily obtained in
perovskite solar cells.17 Meanwhile, to achieve a similar efficiency, a much larger thickness of 3
to 8 m is usually desired for the common thin-film CdTe solar cells.1 This is due to the
5

Chapter 1

ultrahigh optical absorption coefficients of the hybrid perovskite materials, compared to the
traditional thin-film absorbers. According to the density function theory (DFT) calculation, 18 the
lower conduction band of those inorganic absorbers is composed of dispersive s bands while the
bottom of the conduction band in perovskite mainly consists of unoccupied, less dispersive p
bands. By coupling with the p orbital located at the upper valence band, which is provided by the
halogen atom, a strong p-p optical transition becomes possible in perovskite materials.18, 19 As a
result, the direct p-p optical transition in perovskite material gives a higher absorption efficiency
than the CdTe, CIGS absorbers that possess a direct s-p transition. Moreover, the bandgap (Eg) of
perovskite material is around 1.55 eV (800 nm) and to our best knowledge, the highest opencircuit voltage achieved in perovskite cells is 1.1 V, indicating a small deficit between the
bandgap potential (Eg/q) and the Voc (450 meV).8 The high Voc values of perovskite could be
attributed to the lacking of deep trap state within the bandgap, which acts as the non-radiative
recombination centers and shortens the photon-carrier lifetime.20 The formation energies of deep
transition states (Schottky defects) are relatively high while the Frenkel defects with lower
formation energies (such as Pb2+ vacancies and interstitial MA+) only create shallow levels near
band edges,11 as is evidenced by the slight red-shift in the emission spectra of
photoluminescence.21 Besides, unlike inorganic solar cells, grain boundaries and surfaces of
perovskite do not generate deep energy levels in the bandgap. Therefore, the passivation of
surface/grain boundaries is no longer necessary. The reduction of non-radiative charge
recombination centers also suggests a longer carrier lifetime and diffusion lengths in the
perovskite absorber. Because of the low exciton binding energy of 37 meV,22 thermal dissociating
of excitons into unbounded free carriers is highly efficient, which makes it possible for making
planar perovskite solar cells. In conclusion, combined with all these unique properties, perovskite
solar cell has a very promising prospect in PV industry.

1.3.2 Evolution of perovskite device structure


In possessing of these unique photovoltaic properties in perovskite films, considerable efforts
have been paid to increase the power conversion efficiencies of perovskite devices, by using
different device structures and contact layers for the efficient charge collection.
In the beginning, perovskite materials were introduced as the replacement of light sensitizers for
the liquid dye-sensitized solar cells (DSSCs), to improve the light absorption efficiency of
traditional organic sensitizers and reduce the device thicknesses. In DSSCs, a mesoporous layer
of n-type TiO2 is deposited on top of the FTO/compact TiO2 substrate and sensitized with nanocrystalline CH3NH3PbX3 (X = Br, I) particles via a self-organization process, after which the
porous layer is filled with iodide-triiodide redox active electrolyte for the hole transportation. In
2009, the CH3NH3PbI3 sensitized solar cell displayed an efficiency of 3.81%, with a very
promising Voc of 0.96 V.10 Soon the efficiency of perovskite QD-sensitized liquid solar cell was
improved to 6.54%, but the perovskite layer was found to dissolve easily in the liquid electrolyte,
which brought the efficiency down within a few minutes.23 In the meantime, devices using a
solid-state hole transporting material (such as Spiro-OMeTAD) were investigated, with a
stabilized efficiency of 7.6%.24 Furthermore, to find the proof of the efficient charge transport
ability of perovskite layer, instead of using the n-type TiO2, perovskite was deposited onto an
insulating porous scaffold Al2O3.24 It was found that the device with the insulating Al2O3 scaffold
displays the same Jsc and an even higher efficiency of 10.9%, in comparison to the device
assembled via n-type TiO2. Therefore, the perovskite material itself might possess efficient
electron transporting ability. It is then suggested that the main role of the mesostructured
6

Introduction

inorganic oxide scaffold is to provide a larger surface area for heterogeneous nucleation and
crystallization of the perovskite phase. As a result, the formation of small perovskite crystallites
ensures a lower concentration of defects for non-radiative charge recombination.25 Mesoscopic
structure with either active or passive scaffolds are denominated as the mesostructured solar
cells, where initially the superstructure oxide scaffold is covered by a thin perovskite layer, and
the pores are filled with charge transporting material, as is illustrated in Fig. 1.6(a). Also, for the
active scaffolds devices, the possibility of incomplete perovskite covering might induce the
extra recombination loss, which originates from the direct contact between TiO 2 and spiroOMeTAD.26 Since the perovskite material act as charge transporter and light absorber at the same
time, a thick perovskite capping layer is introduced to ensure a complete pore filling and to
enhance the light absorption efficiency, as is shown in Fig. 1.6(b). Mesoporous perovskite solar
cells have achieved an impressive progress in power conversion efficiency, by achieving a PCE
of 20.1% recently.27, 28

Figure 1.625 Mesostructured perovskite device (a) without perovskite capping layer and (b) with perovskite capping
layer. Planar perovskite device in (c) an inverted n-i-p configuration and (d) a regular p-i-n configuration.

Although the mesoporous structures have shown a substantial improvement in efficiencies, the
fabrication of such a complex, well-defined structure is yet a demanding work. In most cases, a
solid-state sintering procedure for the mesoporous scaffold (ca. 450 C) is essential to achieve the
desired device performances.27, 29 The complexity of the fabrication process limits the potential of
large-scale production and therefore, mesoscopic devices only remains interesting for laboratory
studies. As is mentioned above, the successful employment of the insulating oxide scaffold as
well as the capping perovskite layer manifests the ambipolar carrier transport property of
perovskite material, which makes it possible to fabricate planar perovskite devices. To ensure an
efficient charge collection and avoid non-radiative charge recombination, similar to mesoscopic
devices, flat and compact blocking layers are employed for a selective electron collection
(electron transporting layer, ETL) and hole collection (hole transporting layer, HTL),30 while a
perovskite film that acts as both a light absorber and carrier transporter is sandwiched between
them, as is illustrated in Fig. 1.6(c) (inverted configuration n-i-p, where light illuminate from the
n side31) and (d) (regular configuration p-i-n, illuminated from the p side32). Without the
7

Chapter 1

mesoporous scaffold, low-temperature processing (T < 150 C) of perovskite devices becomes


possible and provides a potential application in flexible solar products as well as tandem solar
cells.25 The simplification of the planar perovskite configuration paves the path for different kinds
of deposition techniques and has attained to efficiencies higher than 15%, which is comparable to
many mesoscopic perovskite solar cells.33-35 In this study, research is focused on planar
perovskite solar cells with a regular p-i-n device configuration, where the perovskite layer is
deposited in between the p-type PEDOT:PSS covered Glass/ITO substrate and the fullerene
electron acceptor [60]PCBM, after which LiF/Al is deposited as the counter-electrode. The device
structure and correlated energy diagram are drawn in Fig. 1.7(a) and (b), respectively.

Figure 1.7 (a) Perovskite device structure used in this study. (b) Energy levels of all stacking layers at Voc. Blue arrows
represent the desired charge transfer process; red arrows represent possible charge recombination processes in the
device.

The charge transfer mechanism of perovskite device is shown in Fig. 1.7(b). Under illumination,
photons with energies equal to or higher than the band gap of Eg (1.55 eV) can be absorbed by the
perovskite layer and results in the formation of weakly bound electron-hole pairs. Due to the low
exciton binding energy in perovskite material (37-50 meV), excitons can easily dissociate into
free carriers by thermal energy (kT), where electrons stabilized at the bottom of the conduction
band and holes on top of the valence band. As is shown in Fig. 1.7(b), there are several possible
charge transfer processes in the perovskite layer, among which the injection of electrons from the
perovskite absorber to [60]PCBM and Al electrode (2), the transfer of holes from the perovskite
film to PEDOT:PSS layer and ITO electrode (3) are the desired photon-transition processes that
contribute to an output electrical power. Meanwhile, undesired processes such as the unavoidable
radiative charge recombination (4), non-radiative charge recombination within the perovskite film
(5), recombination of carriers at interfaces of the perovskite and PEDOT:PSS (6), or the
perovskite and [60]PCBM (7), severely reduce the concentration of free carriers in the device,
and result in a much decreased PV performance. However, it is suggested that 36 the timescales of
the process (6)-(7) are much larger than the process of (2) and (3), which ensures an appreciable
charge separation and collection in the planar perovskite device.

1.3.3 Characterization of perovskite solar cells


The photovoltaic performance of perovskite solar cells is typically characterized by measuring the
current density-voltage (J-V) curves both in dark and under ~1000 W/m2 illumination with a
simulated AM1.5G (air mass 1.5 global) solar spectrum, as shown in Fig. 1.8. 1 AM1.5G solar
spectrum correlates to a condition where the sunlight travels a distance 1.5 times the thickness of
8

Introduction

a cloudless atmosphere, with the solar zenith angle of 48.2. In the dark measurement, perovskite
devices display a typical diode behavior, with a low leakage current at a reverse bias. Under
illumination, a photocurrent is generated. At the short-circuit condition (output V = 0), the built-in
electric field sweep photon-generated carriers to the corresponding electrodes for collection, and
the resultant current density is defined as Jsc. The Jsc can be improved by lowering the bandgap of
the absorber, optimizing the active layer thickness and charge mobility. When the forward bias is
equal to the open-circuit voltage (V = Voc), the built-in electric field is balanced by the applied
field, and the current density output becomes 0. In the formation of ohmic contact, Voc correlates
to the bandgap of the perovskite absorber, as well as the recombination processes which
determine the splitting of quasi-Fermi levels in the absorber.37 In between the open-circuit and
short-circuit conditions, the point on the J-V curve which is defined as the maximum power point,
correlates to the maximum power density output when operating the perovskite device under
illumination: Pmpp = JmppVmpp. Therefore, the ratio between the maximum power density and the
area determined by Jsc and Voc is denoted as fill factor (FF), which determines the shape of the J-V
curve, given by

FF

J mpp Vmpp
J sc Voc

FF is dominated by the series and shunt resistance and determines the charge collection efficiency
of the perovskite device. Finally, the power conversion efficiency (PCE) of a solar cell can be
calculated as follows,

PCE (%)

Pmpp
Pin

J sc Voc FF
Pin

Where Pin is the power density of the incident light, usually ~100 mW cm-2. As a result, the PCE
of a solar cell consists of three crucial PV parameters: short-circuit current density Jsc, opencircuit voltage Voc, and fill factor FF.

Current density J (mA/cm2)

illumination
dark

Vmpp

Voc

Pmpp

Jmpp
Jsc
Voltage (V)

Figure 1.8 J-V curve of a perovskite solar cell, measured both in dark and under illumination.

However, since the spectrum generated by a white halogen lamp and a UV filter differs from the
actual AM1.5G spectrum, the Jsc obtained from the J-V curve is not precise even though the
9

Chapter 1

incident light intensity is set at ~100 mW cm-2. Therefore, a spectrally resolved external quantum
efficiency (EQE) measurement is carried out for a precise evaluation of the Jsc and PCE, at the
standard test condition (25 C). EQE is defined as the ratio of the electrons collected in the
external circuit of a device to the number of incident photons at each wavelength. At 100 mW cm2
, by convolution the EQE with the AM1.5G solar spectrum, a corrected Jsc can be calculated,
according to

J sc EQE ( ) E AM 1.5G ( )

e
d
hc

Where e is the elementary charge, h is Planck constant, c is the speed of light. EAM1.5G()
represents the solar energy at the wavelength of , under AM1.5G spectrum.

1.3.4 Deposition techniques for planar PSCs and current issues


For the planar perovskite solar cells, the morphology of the perovskite film has a significant
impact on the final PV performance. In the absence of a mesoporous scaffold, the fabrication of a
compact and well-crystallized perovskite film is challenging. It is reported that38 a high device
performance can be achieved only when the perovskite film displays a high surface coverage. The
incomplete coverage may bring to a direct contact between HTL and ETL, which results in
shunting paths for a leakage current (for Spiro-OmeTAD and TiO2) and, therefore, a lower opencircuit voltage and fill factor. Besides, the presence of less active material in the light traveling
path reduces the light absorption and parallel to a low short-circuit current density. In the
meantime, to realize an efficient charge collection, the formation of large and uniform perovskite
crystals is highly desired to limit the side-effects of grain boundaries and, therefore, maintain high
charge mobilities.39 The main route to achieving an excellent crystallinity is by changing the
thermal annealing conditions of the perovskite layer (to be discussed later), however, because of
the mass transfer phenomena, an improper annealing regime will induce film shrinkage which
increases the number of pinholes and less film coverage. It is, therefore, crucial to control the
balance between the maximum surface coverage and optimum crystallinity.38, 40 On the other hand,
according to transmission spectra of the Glass/ITO/PEDOT:PSS substrate (from 0 to 90%
transmittance, between 300-400 nm),41 the optical absorption range of perovskite material is quite
narrow (300-800 nm). Although the perovskite material possesses a relatively high absorption
coefficient, usually the light absorption is only saturated for < 500 nm, when > 500 nm the
quantum efficiency is highly dependent on the thickness of the active layer, and theoretically,
thicker films result in a stronger light absorption and a higher Jsc.42 However, experimentally, the
highest current density is always achieved in the thickness range of 300-400 nm,17, 42 which
corresponds to a situation that the defects in the films are within control, and the grain sizes of
perovskite are similar to the thickness. Therefore, non-radiative charge recombination processes
are limited, and they possess sufficient diffusion lengths to move across the perovskite layer, and
to be collected at either interface, yielding in an optimized Jsc and Voc.43 To expand the threshold
thickness of perovskite devices while keeping a low concentration of defects in both the bulk and
interfaces, different fabrication methods are exploited, as is shown in Fig. 1.9.
The most general and low-cost deposition method for perovskite film is the one-step spin-coating
method, which involves the spin coating of either a stoichiometric or nonstoichiometric precursor
solution prepared by dissolving lead halide and organic ammonium halide in a common solvent.44
A post-annealing procedure is required to complete the conversion into perovskite crystallites.
10

Introduction

Due to the distinct chemical nature of the inorganic and organic compounds, it is generally
difficult to find a suitable solvent that has a sufficient solubility for both reagents while at the
same time keeps a good control over the reaction, for which the resultant film often suffer from
coarse morphology and crystal agglomeration.45, 46 The extensive crystallization process induces
not only a high density of defects in the layer but also poor surface coverage, which seriously
limits the device performance. Research has been mainly focused on the post-annealing regime
and the precursor composition, to enhance the controllability of the film formation during onestep spin coating.45 It is suggested that34, 47 by using a mixed PbOAc2 and CH3NH3I (1:3)
precursor solution, the precipitation of perovskite crystals is retarded until the volatile organic
byproduct evaporates during the post-annealing. The final perovskite film displays a much better
surface coverage as well as a smoother morphology, compare with the films prepared from PbI2
and PbCl2 lead salts. Recently, it is reported that a combination of PbOAc2, PbCl2, and CH3NH3I
in the precursor composition has achieved with an efficiency higher than 16%, via the one-step
solution process.33

Figure 1.945 Different deposition techniques for perovskite layers.

A similar one-step process is the dual-source vapor deposition, where CH3NH3I and PbCl2 are coevaporated with controlled rates onto a substrate at 10-5 mbar, to achieve a uniform and fully
covered perovskite film (PCE of 15.4%, reported by Liu and coworkers48). In the absence of
solvent, the deposition process alleviates the rapid crystallization as well as the wettability issues
which often lead to the non-uniform coating and pinhole formation, making this method a
suitable candidate for large-area planar perovskite solar cell fabrication. However, due to the high
vapor pressure of CH3NH3I, it is hard to calibrate the evaporation rate of the organic powders
which makes it difficult to control the molar ratio of CH3NH3I to PbCl2, while keeping the
stoichiometry chemical composition is essential for an optimal perovskite solar cell
performance.49 The growth rate via the co-evaporation technique is still too fast to control.
Besides, the use of a high vacuum and a thermal evaporator is energy-consuming and wastes raw
materials, limiting its potential application in the future.
A promising solution deposition technique for smooth perovskite film is the solution-processed
sequential deposition method, which involves the spin-coating of a lead halide precursor solution
onto a substrate in the first step, followed by either dipping the substrate into a solution of organic
11

Chapter 1

ammonium halide to accomplish the conversion to perovskite,50 or spin-coating an upper thick


layer of MAI on top of the lead precursor film,17 and finalized by a post-annealing to accelerate
the inter-diffusion of the organic component into the inorganic layer. Due to the weak van der
Waals gap between interlayers of the PbX2 film,50, 51 a driving force which originates from the
large difference in lattice energy facilitates the intercalation of organic molecules into the
inorganic crystals and enables the conversion of the lead precursor to perovskite. The sequential
deposition relies on the lead halide film that acts as a template to separate the nucleation and
reaction processes, which greatly enhances the surface coverage and homogeneity of the
perovskite films, in comparison to the inferior film morphology obtained in the one-step solution
process. Although promising, however, the solution-processed sequential deposition method is
not yet successful in depositing planar perovskite films, of which the thickness is often below 300
nm.20, 45, 50 Compared to the mesoscopic device structure, the compact nature of PbX 2 film
severely limits its exposure to CH3NH3I solution (to be discussed in Chapter 2) and, therefore,
leads to an incomplete conversion process. As a result, the residual insulating PbX 2 impairs the
device performance as well as the reproducibility of this method. Usually, a much longer reaction
time between the two reagents is needed to reach a complete conversion. However, the backconversion of perovskite over long reaction periods in solution phase51 gives rise to the surface
roughness and may deteriorate the solar cells due to strong leakage current.
By taking advantages of both the rapid reaction rates between the two reagents and the low
sublimation temperature of CH3NH3I, a new deposition approach named vapor-assisted solution
process (VASP) was introduced by Chen and coworkers,52 which involves the in-situ reaction
between the volatile organic vapor and the solid pre-deposited inorganic precursor film.
Compared to the traditional solution processes, the vapor assisted sequential method provides a
moderate reaction condition and therefore, realizes a better control over the reaction kinetics.
Upon prolonging the reaction time, the absence of solvent media and the effective intercalation of
organic vapor molecules will make sure a smooth and complete conversion of the lead halide
precursor. VASP method offers a vigorous re-organizational crystal growth mode that leads to
grain sizes up to micron scale, and the resultant perovskite film displays a good thermostability
with full surface coverage and uniform grain structures.52, 53 This film features collectively
contribute to the highly reproducible efficiency of 12.1%, with an impressive Jsc of 19.8 mA cm-2,
Voc of 0.924 V and FF of 0.663.48
To develop a large-scale and easy-manipulated vapor deposition method, there are several crucial
instrumental parameters to be considered.53 (1) Source temperature: The temperature of the
organic materials should be high enough to provide sufficient concentration of organic vapor
molecules. However, at too high evaporation rate, vapor molecules presented at the substrate will
induce over-rapid reaction and brings to a blocking effect as well as a large surface roughness. (2)
Substrate temperature: The perovskite conversion is an endothermal reaction,54 when organic
molecules land onto the substrate, a high substrate temperature will increase their interdiffusion
and reaction rates with PbX2, which enhances the size and uniformity of the perovskite crystals.
However, at too high substrate temperatures it favors the reverse reaction in which the perovskite
phase decomposes to the lead halide and methyl ammonium salts.55 Besides, it is unfavorable for
organic vapor molecules to deposit on an over-heated substrate and, therefore, limits the
formation of perovskite film.54 (3) Environmental temperature and pressure: Inside the reactor, an
isothermal environment with high temperature and low pressure is preferred, to accelerate the
organic gas phase diffusion.56 These parameters are correlated with each other which makes it
trivial to fabricate high-quality perovskite devices via VASP method. Instead of using Chens
12

Introduction

VASP method, a similar CVD method that uses two temperature zone furnace is designed to
perform the vapor deposition experiments.39, 57

1.4 Outline
The main objective of this report is to fabricate efficient planar-structured perovskite solar cells
via sequential deposition methods and gain a thorough understanding of the formation processes
of perovskite films. Firstly, it is of importance to produce the lead halide precursor films with
controlled morphologies, since these films act as both the structural framework and the reservoir
of reagents for the conversion. Several precursor films are investigated on the basis of their
compositions and morphologies. Secondly, the impact of these precursor features on the reaction
with CH3NH3I both in solution and vapor phase is examined, and the conversion degree of the
precursors is studied in terms of physical properties (optical absorption, layer thickness) of the
perovskite films as well as the device efficiencies. Thirdly, by changing the reaction conditions
with CH3NH3I (concentrations, time, and temperature), the devices are optimized to attain to
decent efficiencies.
In the last part of the report, attention is paid to further investigate the charge transport and
recombination processes inside the perovskite devices. A useful electrical characterization
technique, EIS (Electrochemical impedance spectroscopy), is introduced and applied to highperformance perovskite solar cells. By analyzing the impedance response to a small sinusoidal
bias perturbation imposed on the device, dynamic events in different timescales can be easily
differentiated and interpreted to ascertain the origin of inferior PV performances.

13

Chapter 1

1.5 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.

28.
29.
30.
31.

M. Creatore and R. A. J. Janssen, Solar cells lecture notes, 2015.


I. Dincer, Renew. Sust. Energ. Rev, 2000, 4, 157-175.
J. L. Sawin, Renewables 2015 global status report, Paris, France, 2015.
G. Masson, S. Orlandi and M. Rekinger, Global market outlook for photovoltaics 20142018, 2014, EPIA.
Best research cell efficiencies, http://www.nrel.gov/ncpv/images/efficiency_chart.jpg.
M. A. Green, Prog. Photovolt: Res. Appl., 2009, 17, 183-189.
G. Li, R. Zhu and Y. Yang, Nat. Photonics, 2012, 6, 153-161.
H. J. Snaith, J. Phys. Chem. Lett., 2013, 4, 3623-3630.
D. Weber, Z. Naturforsch., 1978, 33b, 1443.
A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Soc., 2009, 131, 60506051.
M. A. Green, A. Ho-Baillie and H. J. Snaith, Nat. Photonics, 2014, 8, 506-514.
C. Li, X. Lu, W. Ding, L. Feng, Y. Gao and Z. Guo, Acta Cryst. B, 2008, 64, 702-707.
D. Shen, X. Yu, X. Cai, M. Peng, Y. Ma, X. Su, L. Xiao and D. Zou, J. Mater. Chem. A,
2014, 2, 20454-20461.
P. Docampo, J. M. Ball, M. Darwich, G. E. Eperon and H. J. Snaith, Nat. Commun., 2013,
4, 2761.
C. Roldn-Carmona, O. Malinkiewicz, A. Soriano, G. Mnguez Espallargas, A. Garcia, P.
Reinecke, T. Kroyer, M. I. Dar, M. K. Nazeeruddin and H. J. Bolink, Energy Environ.
Sci., 2014, 7, 994.
D. Liu and T. L. Kelly, Nat. Photonics, 2013, 8, 133-138.
Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang, Y. Gao and J. Huang,
Energy Environ. Sci., 2014, 7, 2619.
W. J. Yin, T. Shi and Y. Yan, Adv. Mater., 2014, 26, 4653-4658.
H. S. Jung and N. G. Park, Small, 2015, 11, 10-25.
T.-B. Song, Q. Chen, H. Zhou, C. Jiang, H.-H. Wang, Y. Yang, Y. Liu, J. You and Y.
Yang, J. Mater. Chem. A, 2015, 3, 9032-9050.
Y. Shao, Z. Xiao, C. Bi, Y. Yuan and J. Huang, Nat. Commun., 2014, 5, 5784.
J. M. Frost, K. T. Butler, F. Brivio, C. H. Hendon, M. van Schilfgaarde and A. Walsh,
Nano Lett., 2014, 14, 2584-2590.
J. H. Im, C. R. Lee, J. W. Lee, S. W. Park and N. G. Park, Nanoscale, 2011, 3, 4088-4093.
M. M. Lee, J. Teuscher, T. Miyasaka, T. N. Murakami and H. J. Snaith, Science, 2012,
338, 643-647.
T. Salim, S. Sun, Y. Abe, A. Krishna, A. C. Grimsdale and Y. M. Lam, J. Mater. Chem.
A, 2015, 3, 8943-8969.
T. Leijtens, B. Lauber, G. E. Eperon, S. D. Stranks and H. J. Snaith, J. Phys. Chem. Lett.,
2014, 5, 1096-1102.
F. Giordano, A. Abate, J. P. Correa Baena, M. Saliba, T. Matsui, S. H. Im, S. M.
Zakeeruddin, M. K. Nazeeruddin, A. Hagfeldt and M. Graetzel, Nat. Commun., 2016, 7,
10379.
W. S. Yang, Noh, J.H., Jeon, N.J., Kim, Y.C., Ryu, S., Seo, J. and Seok, S.I., Science,
2015, 348, 1234.
J. H. Im, I. H. Jang, N. Pellet, M. Gratzel and N. G. Park, Nat. Nanotechnol., 2014, 9,
927-932.
J. Seo, S. Park, Y. Chan Kim, N. J. Jeon, J. H. Noh, S. C. Yoon and S. I. Seok, Energy
Environ. Sci., 2014, 7, 2642.
H. Zhang, J. Mao, H. He, D. Zhang, H. L. Zhu, F. Xie, K. S. Wong, M. Grtzel and W. C.
H. Choy, Adv. Energy Mater., 2015, 5, 1501354-1501364.
14

Introduction

32.
33.
34.

35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.

48.
49.
50.
51.
52.
53.
54.
55.
56.
57.

Y. Li, W. Sun, W. Yan, S. Ye, H. Peng, Z. Liu, Z. Bian and C. Huang, Adv. Funct.
Mater., 2015, 25, 4867-4873.
J. H. Heo, D. H. Song, H. J. Han, S. Y. Kim, J. H. Kim, D. Kim, H. W. Shin, T. K. Ahn,
C. Wolf, T. W. Lee and S. H. Im, Adv. Mater., 2015, 27, 3424-3430.
W. Qiu, T. Merckx, M. Jaysankar, C. Masse de la Huerta, L. Rakocevic, W. Zhang, U. W.
Paetzold, R. Gehlhaar, L. Froyen, J. Poortmans, D. Cheyns, H. J. Snaith and P. Heremans,
Energy Environ. Sci., 2016, 9, 484-489.
H.-S. Ko, J.-W. Lee and N.-G. Park, J. Mater. Chem. A, 2015, 3, 8808-8815.
A. Marchioro, J. Teuscher, D. Friedrich, M. Kunst, R. van de Krol, T. Moehl, M. Grtzel
and J.-E. Moser, Nat. Photonics, 2014, 8, 250-255.
F. Fabregat-Santiago, G. Garcia-Belmonte, I. Mora-Sero and J. Bisquert, Phys. Chem.
Chem. Phys., 2011, 13, 9083-9118.
G. E. Eperon, V. M. Burlakov, P. Docampo, A. Goriely and H. J. Snaith, Adv. Funct.
Mater., 2014, 24, 151-157.
P. Luo, Z. Liu, W. Xia, C. Yuan, J. Cheng and Y. Lu, ACS Appl. Mater. Interfaces, 2015,
7, 2708-2714.
A. Dualeh, N. Ttreault, T. Moehl, P. Gao, M. K. Nazeeruddin and M. Grtzel, Adv.
Funct. Mater., 2014, 24, 3250-3258.
Z. Tan, W. Zhang, D. Qian, C. Cui, Q. Xu, L. Li, S. Li and Y. Li, Phys. Chem. Chem.
Phys., 2012, 14, 14217-14223.
Q. Lin, A. Armin, R. C. R. Nagiri, P. L. Burn and P. Meredith, Nat. Photonics, 2014, 9,
106-112.
Z. Xiao, Q. Dong, C. Bi, Y. Shao, Y. Yuan and J. Huang, Adv. Mater., 2014, 26, 65036509.
S. Sun, T. Salim, N. Mathews, M. Duchamp, C. Boothroyd, G. Xing, T. C. Sum and Y.
M. Lam, Energy Environ. Sci., 2014, 7, 399-407.
L. Zheng, D. Zhang, Y. Ma, Z. Lu, Z. Chen, S. Wang, L. Xiao and Q. Gong, Dalton
Trans., 2015, 44, 10582-10593.
J. Y. Jeng, Y. F. Chiang, M. H. Lee, S. R. Peng, T. F. Guo, P. Chen and T. C. Wen, Adv.
Mater., 2013, 25, 3727-3732.
W. Zhang, M. Saliba, D. T. Moore, S. K. Pathak, M. T. Horantner, T. Stergiopoulos, S. D.
Stranks, G. E. Eperon, J. A. Alexander-Webber, A. Abate, A. Sadhanala, S. Yao, Y. Chen,
R. H. Friend, L. A. Estroff, U. Wiesner and H. J. Snaith, Nat. Commun., 2015, 6, 6142.
M. Liu, M. B. Johnston and H. J. Snaith, Nature, 2013, 501, 395-398.
L. K. Ono, M. R. Leyden, S. Wang and Y. Qi, J. Mater. Chem. A, 2016.
J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M.
Gratzel, Nature, 2013, 499, 316-319.
H. A. Harms, N. Tetreault, N. Pellet, M. Bensimon and M. Gratzel, Faraday Discuss.,
2014, 176, 251-269.
Q. Chen, H. Zhou, Z. Hong, S. Luo, H. S. Duan, H. H. Wang, Y. Liu, G. Li and Y. Yang,
J. Am. Chem. Soc., 2014, 136, 622-625.
H. Zhou, Q. Chen and Y. Yang, MRS Bulletin, 2015, 40, 667-673.
J. Mao, H. Zhang, H. He, H. Lu, F. Xie, D. Zhang, K. S. Wong and W. C. H. Choy, RSC
Adv., 2015, 5, 73760-73766.
Q. Chen, H. Zhou, T. B. Song, S. Luo, Z. Hong, H. S. Duan, L. Dou, Y. Liu and Y. Yang,
Nano Lett., 2014, 14, 4158-4163.
M. R. Leyden, L. K. Ono, S. R. Raga, Y. Kato, S. Wang and Y. Qi, J. Mater. Chem. A,
2014, 2, 18742-18745.
Y. Peng, G. Jing and T. Cui, RSC Adv., 2015, 5, 95847-95853.

15

Chapter 2 Solution-processed sequential deposition for planar


perovskite solar cells

Abstract
Inexpensive, simplified and large-scale solution processed sequential deposition technique was
investigated to make efficient perovskite solar cells, in which the chemical nature, as well as the
morphology of the lead halide precursor films, play a crucial role in controlling the growth of
perovskite films. During the first step deposition, three different precursor films, including the
crystallized PbI2, porous PbI2, and mixed halide (by incorporating PbCl2) were fabricated and
characterized in terms of morphologies and optical properties. The impact of these properties of
on the conversion into final perovskite films during the second-step CH3NH3I deposition was
systematically investigated. After that, the sequential deposition procedures, as well as the post
thermal treatments, were optimized and resulted in decent power conversion efficiencies in the
perovskite solar cells. Under simulated AM 1.5G and 100 mW cm-2 illumination, the device made
from the compact PbI2 film exhibits a power conversion efficiency of 7.52%, while PCEs of
devices from porous PbI2 and mixed halide precursor are 8.95% and 8.04%, respectively.

Solution-processed sequential deposition for planar perovskite solar cells

2.1 Introduction
After several years of intense research, PCEs of inorganic-organic hybrid perovskite solar cells
have been fabulously increased from 4% to 21%.1 Among them, most of the efficiently working
PSCs are derived from the mesoporous structures, in which the porous TiO2 or Al2O3 acts as a
scaffold where the perovskite phase deposits. The utilization of such mesoporous structures
requires high sintering temperatures (T > 450 C) and complicated fabrication process, which is
not possible for future developing in roll-to-roll printing on flexible plastic substrates. Therefore,
it only remains interesting in laboratory-scale investigations. In comparison, planar perovskite
devices are more attractive for practical applications. However, to make efficient PSCs, the major
difficulty without using porous scaffold lies in controlling the morphology of the planar
perovskite films, and accordingly, the device performance is closely related to features such as
surface coverage, layer compactness and roughness.2, 3
As is discussed in Chapter 1, there are several deposition techniques developed to study the
formation mechanisms of planar perovskite films and the corresponding cell performances,
among which one major approach is the solution-based processing method. Solution processing
has the advantages of low-temperature processable, inexpensive and easily scalable.4 While the
film quality obtained from the one-step solution deposition method is usually less satisfactory
because of the extensive crystallization and, as a result, a poor surface coverage; the sequential
method,5, 6 which involves the deposition of lead halide precursor films and followed by reacting
with the organic ammonium halide, significantly improves the uniformity and the coverage of
perovskite films. The metal halide precursor film plays a crucial role in the sequential processing
because it acts as a template to form the final perovskite layer, by which the nucleation and
reaction processes are separately controlled.
In this chapter, research was mainly focused on the fabrication and optimization of perovskite
solar cells via a solution-processed sequential deposition method. Several kinds of lead halide
precursor films with controlled morphologies were fabricated, and their following conversion into
perovskite was systematically investigated. Optimizations of the PSC devices were looked into
details by fine-tuning the spin-coating and annealing procedure. Above all, the correlations
between processing methods, film morphologies, and their resultant device performances were
systematically investigated.

2.2 Crystallized PbI2 precursor film


The most commonly used solution-processed sequential deposition involves the formation of a
compact, continuous lead halide precursor film in the first step, followed by the inter-diffusion of
organic methylammonium halide to obtain a perovskite film. The procedure of is schematically
drawn in Fig. 2.1. Yellow PbI2 powder dissolved in DMF was cast on a PEDOT:PSS covered
Glass/ITO substrate to obtain a crystallized PbI2 film. Then a CH3NH3I solution dissolved in 2propanol was spin-coated on top of the precursor film and followed by a post-annealing process
to finalize the reaction of the two components.

Figure 2.1 Conventional inter-reaction via sequential deposition method.7

17

Chapter 2
The PbI2 films fabricated from cold solution, cold substrates or room temperature drying process
showed a turbid yellow color, indicating a rough nature of the PbI2 film on the substrate.
Meanwhile, the best devices were obtained with a hot (70C), supersaturated PbI2 solution (400
mg/ml) spun at high speeds onto the pre-heated substrate (70C) and dried at 70C for several
minutes. Fig. 2.2 shows the scanning electron microscopy (SEM) images of the annealed and not
annealed PbI2 precursor film, however, no distinct differences are observed, both of which reveal
a flat and compact morphology, with small pinholes left on top.

Figure 2.2 Top-SEM images of PbI2 films dried at 70 C (left) and room temperature (right), respectively.

In the preliminary test, it was found that the drying process of the PbI2 precursor film strongly
influences the final device performance. For the cells obtained from annealed (70C) precursor
films, the averaged short-circuit current density Jsc, open-circuit voltage Voc, fill factor FF and
power conversion efficiency PCE are 10.95 mA cm-2, 0.656 V, 0.565 and 4.06%, respectively. In
comparison, the cell made from non-annealed (dried at room temperature) precursor film shows a
much decreased Jsc of 4.14 mA cm-2, Voc of 0.372 V, FF of 0.462 and PCE of only 0.71%. It is
suggested8 that a leakage current will occur if a rough perovskite layer directly contacts with the
Al electrode. As a result, enormous charge carrier recombination would happen which reduces
the photovoltaic performances. In this case, the proximate cause of the rough perovskite layer is
the rough nature of the non-annealed precursor film. Besides, the poor coverage of the obtained
perovskite film leads to lower light absorption, which explains the decreased Jsc of the device.

2.2.1 Effect of CH3NH3I concentration on the device performance


It is known that the photovoltaic performance of perovskite solar cells is highly dependent on the
thickness of the perovskite layer. Usually, a small photocurrent is generated in a thin perovskite
film (< 200 nm) due to the low light harvesting efficiency, especially for the long wavelength
radiation ( > 500 nm),9 see further discussions in Section 2.3.4. However, when the thickness of
the perovskite film exceeds 400 nm, the short-circuit current density still decreases even when the
light absorption is saturated. Meanwhile, the Voc and FF of the devices also start to drop. It is
because, in most of the solution processing conditions, the roughness of the perovskite film will
increase with thickness. Accordingly, defects such as cavities and grain boundaries which will act
as traps for charge recombination.10 Besides, for the commonly investigated tri-iodide perovskite,
CH3NH3PbI3, the diffusion lengths of generated charge carriers is measured to be ca. 100 nm, 11
which is significantly shorter than the thickness of perovskite films. As a result, considerable
charge carriers generated in the central part of the perovskite film will be less likely collected at
either interface and, therefore, the Jsc of those thick devices will be reasonability low.
To achieve the optimum thickness of perovskite film (ca. 300 nm, suggest by Xiao and coworkers.7), the crystallized PbI2 layer with a fixed thickness of ~140 nm was chosen to fabricate
18

Solution-processed sequential deposition for planar perovskite solar cells

perovskite solar cells. Also, the concentration of the CH3NH3I used in the second-step spin
coating should be precisely controlled to convert all the PbI2 into the stoichiometric perovskite.
Otherwise, a mismatch amount of PbI2 and CH3NH3I will lead to a non-stoichiometric perovskite
compound or an insulating PbI2-rich phase,12 which will deteriorate the reproducibility and
efficiency of the device.
1.2
11

Open-circuit Voltage Voc (V)

(b)

MAI concentration (mg/ml)


25
35
45
50
55

Short-circuit current density Jsc (mA/cm2)

10

1.0

0.8

0.6

6
25

-4

30

35

40

45

50

55

0.4

30

Power conversion efficiency (%)

-8

0.6

0.4

0.0

0.2

0.4

0.6

0.8

1.0

0.2

Voltage (V)
(c) 0.8

35

40

45

50

55

50

55

MAI concentration (mg/ml)


8

0.8

-12
-0.2

25

MAI concentration (mg/ml)


1.0

Fill factor

Current density J (mA/cm2)

(a)

25

30

35

40

45

50

MAI concentration (mg/ml)

55

25

30

35

40

45

MAI concentration (mg/ml)

11.54 mA/cm
2
10.81 mA/cm

0.7
0.6

EQE

0.5
0.4
0.3
0.2
0.1
0.0
300

400

500

600

700

800

Wavelength (nm)

Figure 2.3 Current-voltage curves and photovoltaic parameters of perovskite devices fabricated with different MAI
concentrations, from 25 to 55 mg/ml. (a) J-V curve of the highest-performing device for each MAI concentration,
measured under 100 mW cm-2 light illumination. (b) Jsc (top left), Voc (top right), FF (bottom left) and PCE (bottom
right) as a function of MAI concentration. Filled circles represent average values obtained from four solar cells; error
bars indicate the standard deviation. (c) EQE spectra obtained from 45 mg/ml MAI deposition, () under bias light (530
nm green light, AM1.5G spectrum) and () without bias light.

The influence of CH3NH3I concentration on the J-V characteristics is depicted in Fig. 2.3(a), and
all averaged photovoltaic parameters are summarized in Fig. 2.3(b). It is clear that the mean value
of PCEs shows the same trend as Jsc as a function of CH3NH3I concentration, indicating that the
primary factor that determines the device efficiency is the short-circuit current density. Under
illumination of 100 mW cm-2, Jsc firstly increases from 6.64 to 10.43 mA cm-2 together with the
rise in CH3NH3I concentration from 25 to 45 mg/ml, and then it slowly decreases to 7.26 mA cm2
when CH3NH3I approaches 55 mg/ml (summarized in Table 2.1). It was observed that upon
dropping the CH3NH3I solution, the PbI2 substrate displayed an instant color transition from
yellow to red-brown, which means that the reaction between these compounds was extremely fast
and that a perovskite-like complex was already produced during the spin-coating. The
CH3NH3I presented on of the substrate would quickly react with the surface part of the PbI2 film,
and results in the formation of expanded perovskite crystals (to be discussed later) on top, which
blocks further penetration of the organic solution.13 As a result, a large portion of insulating PbI2
below the surface remains intact, giving a much-lowered light absorption and Jsc, as well as an
inferior PCE. In this study, when the CH3NH3I concentration was lower than 45 mg/ml, even less
19

Chapter 2
PbI2-on-top was converted, and therefore, a relatively low Jsc was achieved. When the CH3NH3I
concentration was beyond 45 mg/ml, however, not only did the conversion of the bottom-PbI2 not
increase but also a small amount of insulating CH3NH3I remained on the perovskite film, which
explains the increase in film thickness while the Jsc was reduced.
The highest PCE of 7.73% is attained with 45 mg/ml CH3NH3I, yielding a Jsc of 10.70 mA cm-2,
Voc of 0.948 V, and fill factor of 0.764, as can be seen in Fig. 2.3(a). The integrated current
density from the bias illuminated external quantum efficiency (EQE) curve is 10.81 mA cm-2,
which closely matches with the Jsc measured from the J-V curve.14 Also, a small difference
between EQE curves with and without bias illumination is observed, which could be attributed to
the good morphology of the obtained perovskite film, shown in Fig. 2.4. The film displays a flat
and compact feature, which consists of particles ranging from 100-300 nm, with a high surface
coverage degree and few pin-holes. As the light intensity increases (with the 530 nm bias green
light), instead of recombining with each other, free charges generated in the perovskite layer
could be effectively collected to the same extent as the lower light illumination condition (without
bias light). This phenomenon is also evidenced by the large values of Voc and FF of the device.
Table 2.1 Photovoltaic performances of devices with various CH3NH3I concentrations, determined by 100 mW cm-2
illumination condition. All the PV parameters are averaged values from four cells.

CH3NH3I
(mg/ml)
25
35
45
50
55

Thickness
(nm)
209.77
213.77
218.49
244.80
246.23

Jsc (mA cm-2)

Voc (V)

FF

PCE (%)

6.64 0.40
7.33 1.04
10.43 0.28
8.36 0.71
7.26 0.71

0.958 0.02
0.930 0.12
0.922 0.02
0.917 0.13
0.782 0.04

0.500 0.04
0.643 0.03
0.741 0.03
0.705 0.18
0.676 0.02

3.18 0.39
4.42 1.05
7.22 0.42
5.36 1.80
3.86 0.60

Figure 2.4 Top-view of perovskite film obtained by spin-coating 45 mg/ml CH3NH3I onto
compact PbI2.

The EQE spectrum shows a typical incident photon-to-current behavior of perovskite solar cells,
with the onset of the absorption at 800 nm, in correspondence with the bandgap of CH3NH3PbI3
(1.55 eV).15 According to the transmission spectrum of ITO/PEDOT:PSS,16 the absorption of
perovskite decreases in between 300 and 400 nm. Besides, due to the short penetration depth of
short wavelength of light in perovskite, electrons generated close to the PEDOT:PSS/Perovskite
interface will have to diffuse across the entire perovskite film before collecting by the electron
transporting layer [60]PCBM. However, most of them are recombined, which cuts off the EQE in
the blue part. The EQE displays a pronounced shoulder at 400-600 nm and a peak value of 52%
at 600 nm but decreases notably between 600 and 750 nm, resulting in a massive loss of Jsc, as
compared to the high EQE of 75-80% achieved at 600-750 nm, reported by Liu and co-workers.6
20

Solution-processed sequential deposition for planar perovskite solar cells

As is mentioned above, the depressed perovskite layer thickness might be the reason for the nonabsorption of photons in the long wavelength section, according to the characteristic absorption
spectrum of perovskite material, even though additional light reflected by the Al back electrode
will slightly enhance the absorption efficiency.8 This assumption is evidenced by the layer
thickness of the best-performing device, which was only 218 nm as is displayed in Table 2.1.
In the precursor film, PbI2 is bonded in a hexagonal crystal lattice which consists of layered edgesharing octahedron crystallites, has a density of 6.16 g/cm3 and unit cell volume of 125.7 3.
After the conversion, PbI2 crystals are expanded in the formation of tetragonal CH3NH3PbI3
crystals with a density of 4.22 g/cm3 and unit cell volume of 249.5 3.17 Theoretically, the
volume expansion from PbI2 to perovskite should be 2.1, which is already experimentally proven
by Liu and Zhang et al.10, 13 Since the precursor thickness is 140 nm, it is hypothesized that the
resultant thickness of the perovskite layer should be ca. 290 nm, which hugely deviates from 218
nm and therefore, can be explained by the incomplete reaction between the PbI2 and CH3NH3I, as
is schematically shown in Fig. 2.5. Besides the blocking issue discussed above, another reason for
the incomplete reaction would lie in the deposition procedure, in which the duration of exposure
(to be discussed in next Section 2.2.2) between the CH3NH3I solution and the PbI2 precursor films
was severely limited. In consequence, only the upper part of the PbI2 gets contact with CH3NH3I
while the precursor beneath the interface is isolated from the CH3NH3I solution.

Figure 2.5 A brief illustration showing the residue PbI2 beneath the contact interface.

2.2.2 Effect of loading time on the PbI2 conversion


Due to the compact nature of the PbI2 film, the infiltration capacity of CH3NH3I is restrained, and
hence, a relatively long reaction time (from several minutes to several hours) in solution is
desired to reach a complete conversion to perovskite. Inspired by the one-step solution approach
where an intermixing solution of lead halide and organic ammonium halide is applied, here a
loading time effect of CH3NH3I is investigated. During the second step, a CH3NH3I solution (100
L, 45 mg/ml) was dropped on the PbI2 film and waited for several seconds (loading time) before
spinning. It was hypothesized that the MAI drops would be able to diffuse physically into the
deeper layer of the PbI2 film after some time, which leads to a better pre-mixing of the two
components. Therefore, a higher conversion of PbI2 to perovskite could be achieved.
Fig. 2.6(a) shows the J-V characteristics of the devices fabricated from different loading times,
measured under 100 mW cm-2 light intensity. The variation trends of all the photovoltaic
parameters are summarized in Fig. 2.6(b) and Table 2.2. The averaged Jsc of the devices only
slightly increases from 10.43 mA cm-2 to 10.87 mA cm-2 when the loading time extends from 0 to
10 s, but then steadily decreases with the increasing loading time, and finally, attain to a Jsc of
3.84 mA cm-2 at 60 s. In comparison, when the loading time increases from 0 s to 30 s, both Voc
and FF of the devices maintain good results (Voc 0.9 V, FF > 0.72) without noticeable reduction,
and they only become worse at 60 s loading time. It is clear that the device efficiencies are
basically influenced by the current density (except for the 10 s), with the highest PCE of 7.22%
(reference device, taken from Section 2.2.1) achieved at 0 s loading time, which gradually
decreases from the 6.72% at 10 s, to only 1.23% of 60 s loading time.
21

Chapter 2
Table 2.2 Photovoltaic performances of devices with various loading times, determined at 100 mW cm-2 illumination
condition. All the PV parameters are averaged values from four cells.

0
10
15
30
60

Rq (nm)

Jsc (mA cm-2)

Voc (V)

FF

2.35
2.57
1.93
2.79
18.96

2.89
3.53
2.42
3.43
22.84

10.425
10.867
8.548
5.053
3.835

0.922
0.854
0.928
0.915
0.688

0.741
0.724
0.729
0.725
0.466

(b)

Loading time
0 sec.
10 sec.
15 sec.
30 sec.
60 sec.

PCE
(%)
7.22
6.72
5.77
3.31
1.23

1.1
11
1.0

10

Current density J (mA/cm2)

Ra (nm)

9
0.9
8
7
6

0.8

0.7

5
0.6
4

-4

10

20

30

40

50

60

0.5

10

20

Loading time (s)

0.8

30

40

50

60

40

50

60

Loading time (s)


8

0.7

-8

-12
-0.2

MPP (%)

0.6

Fill factor

Current density J (mA/cm2)

(a)

Thickness
(nm)
214.49
239.42
261.98
238.97
273.36

Voltage (V)

Loading time (s)

0.5

0.4

0.0

0.2

0.4

0.6

0.8

1.0

0.3

0
0

Voltage (V)

10

20

30

40

50

60

10

20

30

Loading time (s)

Loading time (s)

(c)
0.6

0s
10 s
15 s

0.5

EQE

0.4

0.3

0.2

0.1

0.0
300

400

500

600

700

800

Wavelength (nm)

Figure 2.6 Photovoltaic performance of perovskite devices fabricated with different loading times. (a) J-V curves of
the highest-performing device for each loading time (from 0 to 60 seconds), measured under 100 mW cm-2 light
illumination. (b) Jsc (upper left), Voc (upper right), FF (bottom left) and PCE (bottom right) as a function of loading
time. Filled circle represents average values obtained from four solar cells. Error bars indicates the standard deviation.
(c) Biased EQE spectra obtained at 0, 10 and 15 s of loading time.

Interestingly, according to Table 2.2, all the perovskite films with longer loading times (> 0 s)
display higher thickness than the device without loading time (0 s). According to the conclusion
drawn in Section 2.2.1, this increment in thickness with longer loading time could be attributed to
a higher level of completion in the conversion of PbI2 to perovskite, for which an improvement in
Jsc, as well as PCE of the devices, would have been expected. Beyond that, as can be seen in Fig.
2.6(c), for 0, 10 and 15 s loading time, the EQE in the long wavelength range (600 to 780 nm)
slightly increases, which also proves the formation of thicker perovskite films with longer loading
times. However, as is shown above, the photocurrent together with the PCEs shows a decreasing
trend with loading times, despite there was the possibility of a higher conversion reaction. During
the loading treatment, there was a risk of forming a non-stoichiometric perovskite compound,
since the amount of MAI pre-mixed with the precursor PbI2 films was enormous. Cheng and co22

Solution-processed sequential deposition for planar perovskite solar cells

workers18 demonstrated the inhomogeneous composition profile of those loading time treated
perovskite devices, and they also proved the existence of evenly distributed uncoordinated MA +
ions in the bulk, which will act as deep traps for charge recombination. This explains the rationale
in low Jsc as well as PCE at long time loading ( 15 s). Moreover, upon further increasing the
loading time, the film becomes very rough, as is evidenced by the high film roughness at 60 s
loading (see Table 2.2). This was mainly due to the back-conversion of perovskite to PbI2 upon
long time exposure in 2-propanol solution,17 which corresponds to the significant decrease in Voc,
FF, and PCE at 60 s of loading time.
For crystallized PbI2 precursor film, so far, though, better photovoltaic performances have not
been achieved by changing the loading time, it is still meaningful to do more investigation on this
method, given its advantages of simplicity and feasibility of higher conversion reaction.

2.2.3 Effect of annealing time on device performance


For the sequential spin-coating approach, the performance of planar perovskite solar cells is
closely related to the morphology of precursor films, as well as the concentration ratio, and
reaction time involved between the two reactants. Beyond that, the formation of crystallized
perovskite films needs a certain amount of energy,19 for which the energy provided by simply
removing the solvent from substrates is not enough. Therefore, a post-annealing treatment after
spin-coating the organic ammonium halide solution is desired to accelerate the interdiffusion and
reaction between the organic molecules and the inorganic precursor. It is noticeable that at high
annealing temperatures, a rapid crystallization rate will induce large perovskite crystals, however
in the meantime, the pinhole size within the perovskite film also becomes larger, which might
cause a reduction in the surface coverage or compactness. Accordingly, the photocurrent Jsc
decreases because of lowered light absorption, and large gaps between perovskite crystals might
induce charge recombination which deteriorates both Voc and FF. As a result, to control the
balance between the optimum crystallinity and sufficient surface coverage (a small pinhole size),
a moderate post-annealing temperature at 90-100 C is applied to the perovskite film after the
solution deposition process.19, 20
A series of PSC devices fabricated by annealing at 100 C and their photovoltaic performances as
a function of annealing time are summarized in Fig. 2.7. With the increase in annealing time from
15 to 90 minutes, the Voc displays a monotonic decreasing while the Jsc showed an arbitrary
variation. As a result, the highest average PCE of 7.91% is achieved by annealing the device at
100 C for 30 minutes, with a Jsc of 11.4 mA cm-2, Voc of 0.943 V and FF of 0.75. Likewise, upon
longer annealing, the decrease in open-circuit voltage could be attributed to the formation of
larger sizes of pinholes, driven by the tendency of minimizing the grain boundary energy.20
Therefore, the effect of improving the device performance by an increase in perovskite
crystallinity at longer annealing times is offset by the generation of larger pores inside the
perovskite film, which results in similar Jsc values. Different post-annealing times were also
investigated for the devices using 10 and 15 s loading time, and their averaged power conversion
efficiencies are summarized in Table 2.3. It is clear that the devices fabricated by the shortest
loading time (0 s) still give the highest PCEs.
Table 2.3 Photovoltaic performances of devices after annealing at 100 C for various times (15, 30, 60 and 90 minutes).

100 C annealing
0 sec.
10 sec.
15 sec.

15 min.
7.55 %
7.28 %
6.23 %

30 min.
7.91 %
6.82 %
4.99 %

23

60 min.
7.22 %
6.72 %
5.77 %

90 min.
5.74 %
6.86 %
6.49 %

13

1.1

12

1.0

Open-circuit Voltage Voc (V)

Short-circuit current density Jsc (mA/cm2)

Chapter 2

11

10

0.9

0.8

0.7

0.6
20

40

60

80

100

20

Annealing time (min.)

60

80

100

80

100

Power conversion efficiency (%)

0.8

Fill Factor

40

Annealing time (min.)

0.9

0.7

0.6

0.5

5
20

40

60

80

100

20

Annealing time (min.)

40

60

Annealing time (min.)

Figure 2.7 Photovoltaic parameters of perovskite devices (0 s loading time) as a function of annealing time, at 100 C.

2.2.4 Optimized device from c-PbI2 film


In summary, the best device achieved from the crystallized PbI2 is by sequentially spin-coating a
400 mg/ml PbI2 and 45 mg/ml CH3NH3I solution, with 0 s loading time and post-annealing at
100 C for 30 minutes. The active cell area was 0.09 cm2, and a black mask of 0.0676 cm2 was
covered on the device for a more accurate J-V measurement. As can be seen from Fig. 2.8, under
100 mW cm-2 light illumination, a PCE of 7.95% was achieved in the fast, downward (from Voc to
Jsc direction) J-V measurement, yielding a Jsc of 11.4 mA cm-2, Voc of 0.962 V, and FF of 0.725.
In the slow J-V measurement, a slightly higher PCE of 8.18% was achieved, with a Jsc of 11.3
mA cm-2, Voc of 0.93 V, and FF of 0.778. The hysteresis between fast and slow J-V measurements
is low, in corresponding to the optimized perovskite morphology and device fabrication process.
According to the biased EQE spectra, the integrated short-circuit current is 10.4 mA cm-2 and
results in a corrected PCE of 7.52%. Since the solar simulator used for the J-V measurements
does not completely match the AM1.5G spectrum, a slight deviation of the Jsc measured from the
J-V curve and biased EQE spectrum is commonly seen in the results.14
(a) 4

(b) 0.7
2

EQE: 11.07 mA/cm


2
Biased EQE: 10.4 mA/cm

0.6

0.5
-4

EQE

Current density J (mA/cm2)

Fast sweep
Slow sweep

0.4
0.3

-8

0.2
0.1

-12

0.0

0.2

0.4

0.6

0.8

0.0
300

1.0

Voltage (V)

400

500
600
Wavelength (nm)

700

800

Figure 2.8 J-V curves and EQE for the best-performing device produced from porous PbI2. (a) Under 100 mW cm-2
illumination, measurements were taken at fast (0.1 V/s) and slow (0.0013 V/s) sweeps, respectively. (b) EQE spectra,
() under bias light (530 nm green light, AM1.5G spectrum) and () without bias light.

Compare to Xiao7 and Chengs18 best devices in which they employed a similar planar perovskite
structure and two-step solution processing, a decent and comparative Voc (0.962 V) and FF
(0.718), are obtained in this study, indicating the success in optimizing the perovskite film
24

Solution-processed sequential deposition for planar perovskite solar cells

morphology. However, the thickness of the best cell is ~232 nm, significantly lower than the 300
nm of Xiaos and 270~290 nm of Chengs. This explains the low photo-generated current of
11.30 mA cm-2 and, as a result, a relatively low PCE of 8.18% (in the slow scan), in comparison
to Xiao and Chengs results, 19.6 mA cm-2 and 15.4%, 21.82 mA cm-2 and 15.6%, respectively.
To achieve a complete conversion reaction of the PbI2 precursor film, issues such as substrates
wetting, controlled crystallization process of perovskite during the second step, are to be
investigated in the future.

2.3 Porous nano-crystallized PbI2 film


As is discussed in Section 2.2, for planar perovskite solar cells fabricated by the two-step solution
sequential deposition, the biggest obstacle to making good performing devices is the limited
penetration capacity of the organic compound to diffuse into the inorganic PbI2 film. In general,
PbI2 precursor obtained from DMF solution is compact and densely crystallized, which isolates
the MAI solution away from the bottom part of the PbI2 layer, resulting in an incomplete reaction.
To reach a higher conversion, PbI2 films are exposed to the methylammonium halide solution for
a relatively long time, which often leads to the dissolution of perovskite films in the presence of
2-propanol and water,17, 21 evidenced by the worsening of PV performance after long time loading
of MAI solution. In comparison, by using the same deposition procedure, the mesoporous devices
in which a high loading of PbI2 is deposited between TiO2 nanopores, displays a much better
infiltrating ability, where the reaction can be finished within few seconds.5 This is due to the
essence of the porous structure that facilitates the solution penetration. Besides, the confined sizes
of PbI2 will significantly enlarge the surface area and reaction rate of PbI2 crystals. Inspired by
this, a new, porous (p-) PbI2 precursor film was investigated,22, 23 where a small amount of 4-tertbutylpyridine (TBP) solvent was added into DMF to dissolve PbI2. Due to the stronger electrondonating effect of nitrogen atom from TBP than the oxygen atom from DMF, after spin-coating at
room temperature, a film with compound PbI2xTBP was formed after drying the DMF solvent,
and it was not until 30 minutes drying procedure at 70C that the complex started to decompose
to PbI2. Then the sites where TBP molecules resided adjacent to PbI2 became tiny holes, resulted
in a porous structure.

2.3.1 Morphology of p-PbI2 films


A series of precursor solution of 400 mg PbI2, dissolved in 1 mL DMF and various amount of
TBP additive was spin coated on Glass/ITO/PEDOT:PSS substrates, and followed by the drying
procedure to obtain desired porous morphology. Fig. 2.9(a)-(e) shows the SEM images of
precursors made from varied TBP concentrations, which look very different from the pure PbI2
film. By increasing the TBP concentration in the precursor solution, the number of small holes in
the films increases, indicating a porous nature. However, all the films display a flat and
homogeneous feature (especially for Fig. 2.9(c)), where the PbI2 crystals interconnect around the
holes, without forming large aggregations. From the UV-vis absorption measurements, as shown
in Fig. 2.9(f), the obtained precursor films display characteristic PbI2 absorption spectra, with an
absorption onset at around 520 nm. When TBP increased from 0 to 360 L, the sharpness and the
onset position of the absorption edge keep constant, but the peak value of the absorbance started
to decrease. This manifests a higher transmission of incident light caused by a more porous
precursor film, while the nature of the crystallized PbI2 remained unchanged, in corresponded to a
full evaporation of TBP additive after the drying process.

25

Chapter 2

Absorbance (a.u.)

(f) 1.5
0 L
40 L
80 L
120 L
240 L
360 L

1.0

0.5

0.0
400

500

600

700

800

Wavelength (nm)

Figure 2.9 Top-view SEM images of PbI2 precursor films with TBP concentration of (a) 40 L, (b) 80 L, (c)
120 L, (d) 240 L and (e) 360 L. (f) Absorption spectra of precursor PbI2 films with varying TBP
concentrations.

2.3.2 Reaction rate between p-PbI2 films with CH3NH3I


The PbI2120 L TBP precursor film was used to further investigate the conversion into
perovskite film, by dipping the substrate into a CH3NH3I solution of 10 mg/ml in 2-propanol for
different times, followed by rinsing with a 2-propanol solvent to remove excess MAI, and a postannealing process to achieve the final crystallinity. Fig. 2.10(a) shows the effect of dipping time
on the evolution of UV-vis absorption spectra. It is surprising that after only 2 s dipping, the
substrate displays a specific perovskite absorbance near 800 nm while the absorption transition
that originated from the PbI2 (at 500 nm) disappeared. This indicates a quick conversion from the
PbI2 precursor into the perovskite CH3NH3PbI3. Upon dipping the precursor for 20 s, the
26

Solution-processed sequential deposition for planar perovskite solar cells

characteristic absorbance was shown to be the highest, which only slightly decreases when further
extends the dipping time to 10, 20 minutes, indicating a high stability of the obtained perovskite
film. Due to calibration issues, at wavelength > 800 nm, not all the absorption spectra started
from 0 absorbance (sample deposited on glass substrates). Therefore, the difference in absorbance
between 750 and 800 nm (around the absorption edge) versus dipping time was calculated, shown
in Fig. 2.10(b). Still, it is illustrated that the porous PbI2 could quickly react with CH3NH3I,
during which the increase in absorbance is the highest at 20 s (0.26), and then slightly decrease to
0.19 upon 20 minutes dipping.
(b) 0.30

(a) 3.0

2s
20 s
1 min.
2 min.
5 min.
10 min.
15 min.
20 min.
PbI2 precursor

2.0

Absorbance750 nm - Absorbance800 nm

0.25

0.20

Absorbance

Absorbance (a.u.)

2.5

1.5

1.0

0.15

0.10

0.05
0.5
0.00
0.0
400

500

600

700

800

200

400

600

800

1000

1200

Dipping time (seconds)

Wavelength (nm)

Figure 2.10 Effect of dipping time on the evolution of absorption spectra. (a) UV-vis spectra before and after dipping
the porous PbI2120L TBP precursor film into MAI solution. (b) The intensity of perovskite absorption transition as a
function of dipping time.

This observation is distinct from the slow conversion of the conventional planar c-PbI2 film.
Usually, it takes a relatively long (at least tens of minutes 13, 24, 25) dipping time to reach a full
conversion of a 150 nm precursor film to perovskite because of the compact nature and the
blocking effect of conventional PbI2. Besides, during the long time dipping experiments, the
dissolution and peeling-off phenomena are often observed, accompanied by the dramatic
decreasing of absorbance in the UV-vis spectra.24 Both the inhibited reaction and dissolution of
perovskite issues in solution-processed sequential deposition method will seriously limit the
performance and reproducibility of devices.

Figure 2.11 Top-view SEM images of perovskite films by dipping a PbI 2120 L TBP precursor film in a 10 mg/ml
MAI solution for (a) 20 s, (b) 60 s, (c) 120 s, (d) 300 s and (e) 600 s, respectively.

27

Chapter 2
To further investigate the effect of dipping time on the morphology evolution, top-view SEM
images of perovskite films are shown in Fig. 2.11. Upon 20 s dipping, the resultant perovskite
film with a thickness of 262 nm shows a very flat and smooth morphology, which is an indication
of a fast, efficient and yet homogeneous conversion process. As the dipping time increased to 10
minutes, needle-like nanostructures are formed on top of the films. However, the film roughness
and thickness remain almost the same. This is perhaps caused by the slight dissolution and backconversion of perovskite crystals upon prolonged reaction time and followed by a recrystallization process which leads to a less regular shape of the crystal.17

2.3.3 Effect of porosity on the reaction rate


According to the absorption spectra of p-PbI2, as shown in Fig. 2.9(f), higher porosity in the PbI2
film can be achieved by simply adding a larger amount of TBP in the precursor solution. In this
study, to address the differences lay in the reaction rate of the c- and p-PbI2, different PbI2xTBP
precursor films with x varying from 0-360 L were deposited and then dipped into 10 mg/ml
MAI 2-propanol solution for 20 s. Fig. 2.12(a) shows the absorption spectra of perovskite films
obtained from different TBP precursors. Upon 20 s dipping, the film without TBP additive shows
an indistinct onset at 800 nm, which corresponds to the c-PbI2 film that severely blocks the
infiltration of the organic compound. As a result, only the surface part of PbI2 gets reacted, while
the deeper part of PbI2 remains unreacted in the film, as is characterized by the absorption edge at
520 nm. For the 40 L TBP precursor film, after 20 s dipping, the PbI2 absorption edge still exists
but becomes less clear, while the onset of perovskite absorption at 800 nm becomes sharper, both
of which reveal that the higher conversion reaction has been achieved via a more porous PbI2 film.
Above 80 L, the PbI2xTBP precursor films are found to convert to perovskite film much more
quickly within 20 s dipping time, by showing a sharp perovskite absorption onset at ~800 nm,
while no PbI2 absorption edge is observed. Again, to evaluate the sharpness of the characteristic
perovskite absorption, the difference in absorbance between 750 nm and 800 nm is shown in Fig.
2.12(b). It is demonstrated that, after 20 s reaction time, the increment in perovskite absorbance
dramatically increases from 0.08 for 0 L, to 0.26 for 120 L, and then decreases to 0.15 for 360
L.
(a) 3.0

(b)

0.24

0.20

Absorbance

2.0

Absorbance (a.u.)

Absorbance750 nm - Absorbance800 nm

0 L
40 L
80 L
120 L
240 L
360 L

2.5

1.5

1.0

0.16

0.12

0.08

0.5

0.0
400

0.28

0.04
450

500

550

600

650

700

750

800

40

80

120

160

200

240

280

320

360

TBP concentration (L)

Wavelength (nm)

Figure 2.12 Effect of TBP concentration on the evolution of absorption spectra. (a) UV-vis absorption spectra for
PbI2x L TBP with 20 s dipping time, where x is 0, 40, 80, 120, 240 and 360. (b) The intensity of perovskite
absorption transition as a function of dipping time.

As is mentioned above, more TBP additives in the precursor solution result in the higher PbI2
porosity, which facilitates an efficient capillary penetration of the CH3NH3I solution. Also, the
enlarged surface area of p-PbI2 crystals provides more active sites for the reaction with MAI
molecules. Therefore, by increasing the concentration of TBP (but, no larger than 120 L), a
higher and quicker conversion of PbI2 into CH3NH3PbI3 is accomplished, which explains the
28

Solution-processed sequential deposition for planar perovskite solar cells

increase in the UV-vis absorption. When the concentration of TBP surpassed 120 L, however, a
precursor film with too much free space in between the PbI2 crystals was obtained, as is
evidenced by the continuous decreasing in PbI2 absorption at higher TBP concentrations, as
shown in Fig. 2.9. After the conversion with MAI, the resultant perovskite film could also be
highly porous,23 leading to a much decreased light-harvesting efficiency, even though in this case
the conversion of PbI2 to perovskite could be higher.

Figure 2.13 Top-view SEM images of perovskite films by dipping the PbI2x L TBP precursor (a) x = 40 L, (b) x =
120 L, and (c) x = 360 L into a 10 mg/ml CH3NH3I solution for 20 s.

Fig. 2.13 shows the top-view SEM images of perovskite films converted from different TBP
precursors. It is clear that, after 20 s dipping in a CH3NH3I solution, all the films show a similar
uniform and smooth morphology. However, due to the low magnification of these images, less
information about the film compactness can be obtained.

2.3.4 Photovoltaic performance of devices made of PbI2120 L TBP precursor


The optimized PbI2120 L TBP precursor film mentioned above was used to convert to
CH3NH3PbI3, after which the planar devices with the regular structure of Glass/ITO/PEDOT:
PSS/CH3NH3PbI3/[60]PCBM/Al were fabricated. Fig. 2.14 demonstrates that both the drying
time of the p-PbI2 as well as the post-annealing regime for the perovskite film strongly influence
the photovoltaic performance of devices. All the J-V curves were carried out under 100 mW cm-2
light illumination, and the PV parameters are summarized in Table 2.4.
Table 2.4 Photovoltaic performances of devices with varied drying and annealing regimes, determined by 100 mW cm2 illumination condition. All the PV parameters are averaged values from four cells.
Drying & Post-annealing
Dry at 70 C 10 min., anneal at RT
Dry at 70 C 10 min., anneal at 70 C 30 min.
Dry at 70 C 10 min., anneal at 100 C 5 min.
Dry at 70 C 10 min., anneal at 100 C 30 min.
Dry at 70 C 30 min., anneal at 100 C 30 min.

Jsc (mA cm-2)


3.76 0.47
1.78 0.43
4.47 0.23
9.20 1.68
12.05 0.19

Voc (V)
0.934 0.01
0.920 0.04
0.942 0.05
1.018 0.01
0.932 0.01

FF
0.515 0.02
0.539 0.04
0.558 0.12
0.552 0.07
0.676 0.01

PCE (%)
1.81 0.26
0.92 0.28
2.38 0.71
5.26 1.59
7.59 0.17

After drying the precursor film at 70C for 10 minutes and post-annealing the perovskite film at
100C for 30 minutes, the average values of Jsc, Voc, FF, and PCE are 9.2 mA cm-2, 1.018 V,
0.552 and 5.26%, respectively. As the drying time increased up to 30 minutes, Jsc and FF rise to
12.05 mA cm-2 and 0.676, respectively, while the Voc decreased to 0.932 V, and a PCE of 7.59%
was obtained. As is discussed before, TBP is a strongly coordinative solvent for PbI2. After a
short time drying, the PbI2-TBP complex still exists in the film due to the low evaporation rate of
TBP molecules, which brings about a less porous PbI2 structure, and results in a less complete
conversion of PbI2 into CH3NH3PbI3. By extending the drying time, a more porous PbI2 precursor
film is obtained, which contributes to a higher conversion to perovskite film and Jsc of the device.
In the meantime, the increased porosity of PbI2 film might also result in a slightly porous
CH3NH3PbI3 film, where the pores accelerate charge recombination and lower the Voc. However,
29

Chapter 2
it is noticeable that if the porosity of the film is within controlled, perovskite crystals will remain
interconnected so as to provide sufficient light absorption and carrier transport, and as a result, a
better device performance can be achieved.
dry
dry
dry
dry
dry

Current density J (mA/cm2)

at
at
at
at
at

70C
70C
70C
70C
70C

10
10
10
10
30

min.,
min.,
min.,
min.,
min.,

no anneal
anneal at 70C 30 min.
anneal at 100C 5 min.
anneal at 100C 30 min.
anneal at 100C 30 min.

-4

-8

-12

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Volatge (V)

Figure 2.14 Effects of drying and post-annealing on the final device performance. J-V curves of the highest-performing
devices of each substrate, under 100 mW cm-2 light illumination.

Furthermore, the annealing regime during the formation of perovskite films also plays a
significant role in determining the final device performance. It is evidenced by the considerable
increase in Jsc and slightly growth in Voc, from 3.76 to 9.2 mA cm-2, and from 0.934 to 1.108 V,
respectively, when the annealing temperature rises from room temperature to 100 C and
annealing time from 5 to 30 minutes. The increase in open-circuit voltage can be attributed to the
enlarged grain size of the perovskite crystals. As a result, the number of grain boundaries (traps
for recombination) decreases, which improves the carrier transporting ability in the active layer.
(b) 0.8

Fast
Slow

No bias, 14.79 mA/cm


2
Biased, 13.78mA/cm

0.7

0.6
0.5

-4

EQE

Current Density J (mA/cm2)

(a) 4

0.4
0.3

-8

0.2
0.1

-12

0.0

0.2

0.4

0.6

0.8

0.0
300

1.0

Voltage (V)

400

500

600

700

800

Wavelength (nm)

Figure 2.15 J-V curves and EQE for the best-performing device produced from porous PbI2. (a) Under 100 mW cm-2
illumination, measurements were taken at fast (0.1 V/s) and slow (0.0013 V/s) sweep, respectively. (b) EQE spectra, ()
under bias light (530 nm green light, AM1.5G spectrum) and () without bias light.

Fig. 2.15 shows the photovoltaic performance of the best device having been achieved with the
porous precursor structure. In the fast, downward J-V measurement, the device displays a PCE of
7.80%, yielding a Jsc of 12.30 mA cm-2, Voc of 0.938 V and FF of 0.676. In contrast, the slow J-V
measurement gives a slightly higher power conversion efficiency of 8.05%, with a Jsc of 12.42
mA cm-2, Voc of 0.910 V and FF of 0.714. As can be seen from Fig. 2.15(b), the bias dependence
between non-bias (lower light intensity) and bias illuminated EQE (~AM1.5G) curves is low,
which correlates to the smooth perovskite morphology and thus resulting in efficient charge
carrier collection. The integrated Jsc of 13.78 mA cm-2 from biased EQE curve shows a 10%
mismatch compared to the Jsc measured in the J-V curve, which gives a corrected PCE of 8.95%.
30

Solution-processed sequential deposition for planar perovskite solar cells

Also, between 350 and 650 nm, the EQE displays a very high incident photon-to-electron
conversion efficiency, reaching a maximum point of 70% at 510 nm. This is comparable to Zhang
and co-workers observation,22 in which a maximum peak of ~70% was found at 520 nm, by
using a similar precursor recipe and fabrication method. However, the EQE dramatically
decreases between 650 and 780 nm, corresponding to an insufficient perovskite thickness (~262
nm) and, therefore, contributes to a relatively low Jsc of 13.78 mA cm-2, in striking contrast to the
Jsc of 21.27 mA cm-2, reported by Shi and co-workers.23 To reach a better conversion, further
investigations can be performed on increasing the MAI concentrations and fine-tuning the
dipping procedures.
It is interesting to note that there is a valley appearing in the EQE spectra at ca. 580 nm. These
phenomena are also seen in other devices, but the valley is often found to be located at a different
wavelength, in correlation to different device thicknesses. According to Lin and co-workers
calculation,9 when < 500 nm, for any properly functioned perovskite film that has a thickness
larger than 150 nm, light absorption, as well as EQE, is fully saturated, and no optical-field can
reach the back reflective Al electrode. Therefore, the low EQE in the short wavelength range is
mainly caused by transmission loss of the ITO/PEDOT:PSS layer, as well as charge
recombination due to the short lifetime of charge carriers, see discussion in Section 2.2.1. When
> 500 nm, due to the low absorption coefficient, the light will travel through perovskite film
with the little amount being absorbed, after which it is reflected back to the active layer again by
the Al electrode. The reflected light with a certain wavelength might have a destructive
interference with the incoming light, resulting in a much-lowered absorption and a valley at the
specific wavelength in the EQE spectrum.

2.4 Mixed halide perovskite CH3NH3PbI3-xClx film


As previously mentioned, to achieve sufficient light absorption efficiency, the optimized
thickness of planar perovskite layer is around 300~400 nm, according to the penetration depth
profile of sunlight. However, the devices in Section 2.2-2.3 were made of tri-iodide perovskite
CH3NH3PbI3, of which the diffusion length of photon-generated carriers (electrons, holes or
weakly bound excitons) is measured to be only ~100 nm.11 Therefore, for a thick tri-iodide
perovskite film, there is a possibility that charge carriers get recombined before being extracted
by either blocking layer, and resulting in a poor photovoltaic performance. To achieve an efficient
charge collection, a simple method by changing the chemical environment evolved in the
fabrication of perovskite film is suggested, in which chloride is incorporated into the precursor
solutions to form the so-called mixed halide perovskite, CH3NH3PbI3-xClx, with carrier diffusion
lengths larger than 1 m.11, 26-30 By doing so, higher photovoltaic performance becomes
achievable in thick planar perovskite devices. The mechanism for improving the charge
transporting ability is still debatable since the quantified Cl content in the resultant perovskite
film is very low. Also, whether the Cl is incorporated into the crystal lattice of CH 3NH3PbI3 or
simply positioned at the grain boundaries, remains to be investigated.3 Nevertheless, comparing
to their tri-iodide perovskite counterparts, it is commonly observed that the mixed halide
perovskite shows an improved film morphology, which displays the higher surface coverage, the
oriented growth of perovskite crystals, and larger grain sizes.12 Combining all these
morphological characteristics will contribute to a better device performance and reproducibility.

2.4.1 Effect of Cl content on the formation of lead halide precursor films


Similarly, to achieve a better morphology control, the solution-processed sequential deposition
method was chosen to fabricate mixed halide perovskite device, during which the lead halide
precursor film could act as a template for a smooth conversion into perovskite. Besides, it is
reported that the sequential deposition might produce higher Cl concentration in the perovskite
31

Chapter 2
film compare to the one-step deposition.31 In this study, white powders of PbCl2 were used as the
chloride source in the first deposition step. In order to increase the solubility of PbCl2 and ensure
a high Pb2+ concentration for the reaction, a series of mixed PbI2 and PbCl2 in DMF solutions
were made, in which the total concentration of Pb2+ was kept at 1.0 M, while the molar fraction of
PbCl2 changed from 0 to 0.5 ( 0.5, PbCl2 is no longer soluble in DMF), that is, (1-x) M PbI2 + x
M PbCl2 (x = 0, 0.1, 0.2, 0.3, 0.4 and 0.5). As a reference, pure PbCl2 precursor solution of 1 M
PbCl2 in DMSO was also made. However, the PbCl2 did not fully dissolve and therefore, the
solution was filtered before use. All these precursor solutions were then spin-coated on top of the
Glass/ITO/PEDOT:PSS substrates at room temperature, followed by a drying procedure at 70 C
for 30 minutes to obtain the mixed halide precursor films with varied Cl contents, and cooled
down to room temperature, before dipping the substrates into a CH3NH3I solution to convert into
perovskite.
(a)1.5

1.0 M PbI2
0.9 M PbI2 + 0.1 M PbCl2
0.8 M PbI2 + 0.2 M PbCl2
0.7 M PbI2 + 0.3 M PbCl2

Absorbance (a.u.)

0.6 M PbI2 + 0.4 M PbCl2


0.5 M PbI2 + 0.5 M PbCl2

1.0

1.0 M PbCl2 (DMSO)

0.5

0.0
400

500

600

700

800

Wavelength (nm)

Figure 2.16 UV-vis absorption spectra of mixed halide precursor films.

Fig. 2.16 shows the UV-vis absorption spectra of precursor films obtained with different lead
chloride concentrations. As is already explained in Section 2.3.1, the 1.0 M pure crystallized PbI2
film displays a characteristic PbI2 absorption peak at around 520 nm. In comparison, the
precursor film fabricated from 1.0 M pure PbCl2 solution (only partially dissolved in DMSO)
displays almost zero light absorption in the range of 400 to 800 nm, indicating an amorphous
morphology. When the concentration of PbCl2 in the precursor solution increases, it is clear that
the onset position of the PbI2 absorption peak starts to shift slowly to the left, and accordingly the
specific absorption height in this region decreases. After the mole fraction of Cl reaches 0.5, that
is, for the 0.5 M PbI2 and 0.5 M PbCl2 mixed solution, the PbI2 absorption totally disappears. This
variation manifests the formation of a less crystallized intermediate phase between PbI2 and PbCl2.
Wu and co-workers32 reported that by mixing PbI2 with PbCl2 in a 1:1 molar ratio in DMF, a low
crystallinity PbICl phase will be produced. This phenomenon also explains the increase in
solubility of PbCl2 in DMF (at room temperature, the solubility limit of pure PbCl 2 in DMF is
0.29 M32), after incorporating a large amount of PbI2 in the precursor solution.
Fig. 2.17 displays the morphology evolution of lead halide films by adding a different quantity of
PbCl2 in the precursor solution. Compare to the pure PbI2 film (x = 0) which shows small holes
(see Fig. 2.2), when the fraction of PbCl2 in the solution increases to 0.1, the resultant precursor
film becomes more compact and better covered with large and flake-like crystals. However, when
further increase the PbCl2 concentration in the precursor solution (x 0.2), the size of these sheetlike crystals gradually decreases and precursor films with a porous nature are obtained. The
morphology evolution by increasing PbCl2 concentrations in the precursor is consistent with the
observations of others.28, 29 Similar to the porous PbI2TBP precursor film discussed in Section

32

Solution-processed sequential deposition for planar perovskite solar cells

2.3.3, one would have expected that the porous morphology might facilitate the reaction with
CH3NH3I during the second-step deposition.

Figure 2.17 Surface SEM images of mixed halide precursor films. (a) (x = 0.1) 0.9 M PbI2 + 0.1 M PbCl2, (b) (x = 0.2)
0.8 M PbI2 + 0.2 M PbCl2, (c) (x = 0.3) 0.7 M PbI2 + 0.3 M PbCl2 and (d) (x = 0.4) 0.6 M PbI2 + 0.4 M PbCl2.

2.4.2 Conversion of mixed halide precursor into perovskite


To investigate the conversion of mixed halide precursor into mixed halide perovskite film, the
above-mentioned (1-x) M PbI2 + x M PbCl2 (x = 0.1, 0.2, 0.3 and 0.4) precursor substrates were
dipped into a 10 mg/ml CH3NH3I solution for 60 s, followed by a rinsing procedure to remove the
excess CH3NH3I, and a post-annealing procedure at 100 C for several minutes to produce a good
perovskite crystallinity. It is notable that, the CH3NH3I solution was kept at a temperature of
70 C during the dipping procedure because the formation energy needed for Cl to incorporate
into perovskite lattice is higher than that of iodide.31
(b)

(a) 3.0

0.9 M PbI2 + 0.1 M PbCl2 dip 60s

0.24

Absorbance750 nm - Absorbance800 nm

0.8 M PbI2 + 0.2 M PbCl2


0.7 M PbI2 + 0.3 M PbCl2

2.5

0.22

0.6 M PbI2 + 0.4 M PbCl2

2.0

1.0 M PbCl2 (DMSO)

Absorbance

Absorbance (a.u.)

0.5 M PbI2 + 0.5 M PbCl2

1.5

1.0

0.18

0.16

0.14

0.5

0.0
400

0.20

500

600

700

0.12
0.0

800

0.2

0.4

0.6

0.8

1.0

PbCl2 concentration (mol/L)

Wavelength (nm)

Figure 2.18 Effect of chloride content on the evolution of perovskite absorption. (a) Absorption of perovskite films,
after hot dipping (70 C) the precursor films into a 10 mg/ml MAI solution for 60 s. (b) The intensity of perovskite
absorption transition as a function of chloride concentration, between 750 nm and 800 nm.

The absorption spectra of the mixed halide perovskite films are shown in Fig. 2.18(a), and Fig.
2.18(b) displays the transition intensity as a function of PbCl2 concentration. After 60 s dipping, it
is demonstrated that all the films show the characteristic perovskite absorption, with the onset of
33

Chapter 2
absorption edge at around 800 nm. When x increases from 0.1 to 0.4, the calculated sharpness of
the absorbance roughly increases from 0.19 to 0.24, very close to the absorption achieved from
the PbI2120L TBP precursor (0.26), which means that a quick conversion reaction with MAI is
realized by incorporating PbCl2 in the precursor film. However, a distinct PbI2 absorption at 520
nm can still be seen for all the resultant mixed halide perovskite films (x = 0.1, 0.2, 0.3 and 0.4),
indicating that not all the lead in the precursor was converted into perovskite. Followed by the
same procedure, a pure PbCl2 precursor film (1 M in DMSO) was dipped into MAI solution to
achieve the conversion, and its absorption spectrum is also displayed in Fig. 2.18(a). It is very
interesting to observe that the perovskite film made of pure PbCl2 still displays a characteristic
PbI2 absorption peak. Meanwhile, the film gives a high absorption in the long wavelength ( >
800 nm) range, which is caused by the light scattering effect because of the high roughness of the
film. It is suggested that, during the post-annealing process, a possible side reaction can occur in
which the iodide exchange with the chloride in the resultant film, given by Reaction 2.1. 12 As a
result, the volatile organic byproducts methylammonium chloride (MACl) and methylammonium
iodide (MAI) will be evaporated during the annealing procedure and lead to a rougher perovskite
morphology.

3MAI PbCl2 PbI 2 2 MACl MAI

(2.1)

This reaction provides a rational explanation for the emergence of PbI2 in the absorption spectra
of different perovskite films (x = 0.1, 0.2, 0.3, 0.4 and 1.0). Of course, both the limited
penetration of organic solution, as well as the slow conversion of mixed halide precursor, could
be the reasons for the incomplete reaction, which remains further investigations.

2.4.3 Photovoltaic performance of mixed halide perovskite devices


2.4.3.1 J-V performance of different PbCl2 concentration
To elucidate the effect of Cl incorporation on the solar cell performance, devices based on the
planar structure Glass/ITO/PEDOT:PSS/CH3NH3PbI3-xClx/[60]PCBM/Al were fabricated, of
which the CH3NH3PbI3-xClx films were made by hot dipping the mixed halide precursor film (x =
0, 0.1, 0.2, 0.3, and 0.4) into a 10 mg/ml MAI 2-propanol solution for 60 s, and followed by a
post-annealing procedure at 100 C for 30 min. Fig. 2.19(a) demonstrates that under 100 mW cm2
illumination, J-V performance of the perovskite solar cell is strongly influenced by the chloride
content in the precursor solution. All the photovoltaic parameters have been plotted and
summarized in Fig. 2.19(b) and Table 2.5, respectively. It is clear that the average PCE of the
devices shows a similar variation trend with the Jsc and Voc. When the mole fraction x of PbCl2 in
the precursor solution increases from 0 to 0.1 (0.1 M PbCl2), the Jsc rapidly increases from 3.94 to
13.2 mA cm-2, which corresponds to a steep enhancement in the average PCE from 1.29 to 7.36%.
In this case, the device also displays high Voc and FF, of 0.931 V and 0.6, respectively. Compared
to the pure PbI2 (x = 0), the small amount PbCl2 in the precursor film must contribute to a higher
conversion rate into perovskite, as is evidenced by the increase of absorption, and a growth in the
thickness of the resultant perovskite layer, from 208 to 259 nm, shown in Table 2.5. Yu and coworkers33 reported that the formation rate of perovskite film is slowed down due to the MA + rich
environment built by the sequential deposition process. Therefore, the presence of some Cl in the
precursor film helps to remove the excess MA+ by producing volatile MACl vapor which
accelerates the reaction, according to Reaction 2.2. When further increases the PbCl2
concentration to 0.4 M, however, in contrast to its high absorption (see Fig. 2.18(b)) and a
perovskite thickness of ~270 nm, the device performances dramatically decline, yielding a PCE
of only 0.44%, with a Jsc of 3.02 mA cm-2, Voc of 0.501 V and FF of 0.286. This is possibly
related to the extensive evaporation of gaseous by-product MACl during the reaction process
(also some MAI evaporated during a side reaction, see Reaction 2.1), which results in a rough
34

Solution-processed sequential deposition for planar perovskite solar cells

perovskite layer that lowers the photovoltaic performance of the devices, even though in this case
the conversion to perovskite might be higher.

3MAI PbCl2 MAPbI 3 2 MACl

(2.2)

In addition, different annealing procedures for 0.1 M PbCl2 precursor and perovskite films are
investigated, but no obvious trends are found, which might be related to the uncertain conversion
degrees of the lead source.
(b)

1.0 M PbI2

0.9 M PbI2+0.1 M PbCl2

12

Current density J (mA/cm2)

0.8 M PbI2+0.2 M PbCl2


0.6 M PbI2+0.4 M PbCl2

0.8
10

0.6

-4

0.4
0.0

0.1

0.2

0.3

0.4

0.0

0.1

PbCl2 molar fraction

0.2

0.3

0.4

0.3

0.4

PbCl2 molar fraction

0.8

0.7

-8
0.6

-12

PCE (%)

0.5

Fill Factor

Current density J (mA/cm2)

0.7 M PbI2+0.3 M PbCl2

1.0
14

Voltage (V)

(a)

0.4

0.3
0.2

0.1

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

0.0

0
0.0

Voltage (V)

0.1

0.2

0.3

PbCl2 molar fraction

0.4

0.0

0.1

0.2

PbCl2 molar fraction

Figure 2.19 Current-voltage curves and photovoltaic parameters of perovskite devices fabricated with different PbCl 2
ratios, from 0 to 0.4 mol/L. (a) J-V curves of the highest-performing devices for each PbCl2 concentration, measured
under 100 mW cm-2 light illumination. (b) Jsc (top left), Voc (top right), FF (bottom left) and PCE (bottom right) as a
function of PbCl2 concentration. Filled circles represent average values obtained from four solar cells. Error bars
indicates the standard deviation.
Table 2.5 Photovoltaic performances of devices with varied MAI concentrations, determined at 100 mW cm-2
illumination condition. All the PV parameters are averaged values from four cells.

PbCl2 (mol/L)
0
0.1
0.2
0.3
0.4

Thickness (nm)
208.24
258.72
288.84
280.51
274.75

Jsc (mA cm-2)


3.94 0.97
13.20 0.32
11.15 0.30
7.82 0.46
3.02 0.13

Voc (V)
0.930 0.02
0.931 0.02
0.724 0.01
0.545 0.02
0.501 0.02

FF
0.333 0.09
0.600 0.04
0.556 0.04
0.341 0.03
0.286 0.01

PCE (%)
1.29 0.57
7.36 0.41
4.49 0.40
1.46 0.19
0.44 0.03

2.4.3.2 Optimized mixed halide perovskite device


Fig. 2.20 displays the J-V curve and EQE spectra of the best-performing mixed halide device.
Under 100 mW cm-2 light illumination, the device fabricated from 0.9 M PbI2 + 0.1 M PbCl2
precursor shows a PCE of 7.97% in the fast J-V measurement, with Jsc = 13.4 mA cm-2, Voc =
0.897 V and FF = 0.662. After the slow J-V measurement, however, a hysteresis effect between
two J-V curves is observed. Both Voc and FF increase while the Jsc decreases remarkably, yielding
a value of 1.035 V, 0.719 and 10.8 mA cm-2, respectively. As a result, the PCE of 8.04%
presented was comparable to the data obtained from the fast measurement. Considering the slow
measurement procedure in this study, in which the perovskite solar cells were exposed to the 100
mW cm-2 light illumination for 30 minutes under a forward bias voltage (close to Voc), the
enhancement in Voc and FF should be attributed to the light-soaking effect.34 Under prolonged
light illumination and forward bias, substantial electrical dipoles (or accumulated space charges)
with directions opposite to the applied voltage will be built, due to the possible existence of the
ferroelectric polarization in planar perovskite crystals, which promotes the separation of photogenerated carriers and increases the mobility of charges to reach to either blocking layer. The
35

Chapter 2
combination of these effects will certainly improve the Voc of the device. Also, one possible
reason for the severe decrease of Jsc during the slow measurement may relate to the high defect
density in the solar cell. It is suggested that35 under the illumination and an applied forward bias
(all the J-V measurements were taken in the direction from open-circuit to short-circuit), the
original deep traps in the perovskite film are filled up and stabilized by opposite charges.
Therefore, when the fast J-V sweep comes close to the short-circuit point, during which the scan
rate is faster than the de-trapping rate, the charges could quickly transfer to the selective
collection layers without inducing significant trapping processes, in correspondence to a higher
Jsc. In comparison, during the slow scan measurements, a possible de-trapping process and, as a
result, more serious charge recombination would happen, which slows down the charge collection
and cuts down the Jsc.
(a)

(b) 0.8

EQE, 14.92 mA/cm


2
Biased EQE, 13.44 mA/cm

0.7
0.6

0.5
-4

EQE

Current density J (mA/cm2)

Fast, 7.97%
Slow, 8.04%

0.4
0.3

-8

0.2
0.1

-12

0.0

0.2

0.4

0.6

0.8

0.0
300

1.0

Voltage (V)

400

500
600
Wavelength (nm)

700

800

Figure 2.20 J-V curves and EQE for the best-performing device produced from CH3NH3PbI3-xClx. (a) Under 100 mW
cm-2 illumination, measurements were taken at fast (0.1 V/s) and slow (0.0013 V/s), respectively. (b) EQE spectra
under biased and non-biased illumination conditions.

The biased EQE in Fig. 2.20(b) shows an integrated Jsc of 13.44 mA/cm-2, obtained from the best
mixed halide perovskite device (259 nm), and it is compared with the EQE curves of the best
perovskite devices shown in previous sections, as can be seen in Fig. 2.21. It is found that the pPbI2 film provides the most rapid conversion to perovskite (262 nm, 13.78 mA cm-2) while the
compact PbI2 film displays the lowest conversion rate (232 nm, 10.4 mA cm-2).
0.8

0.1 M PbCl2
Porous PbI2
Compact PbI2

EQE

0.6

0.4

0.2

0.0

300

400

500

600

700

800

Wavelength (nm)

Figure 2.21 EQE spectra of () Mixed halide perovskite film made from 0.9 M PbI2 + 0.1 M PbCl2 precursor, () triiodide perovskite film manufactured from the porous PbI2 precursor and () tri-iodide perovskite made from the
compact PbI2 precursor.

On the other hand, no obvious change in the onset of the EQE curve for mixed halide perovskite
is observed. Also, no clear blue-shift is seen in perovskite absorption spectra with chloride
36

Solution-processed sequential deposition for planar perovskite solar cells

incorporation, as shown in Fig. 2.18(a). Both phenomena reveal that the addition of PbCl2 in the
precursor solution does not induce significant bandgap modification, in contrast to Li and coworkers31 observation (However, they used a higher PbCl2 concentration in the precursor).
Therefore, it comes to a conclusion that the main roles of the chloride are to fine-tune the
conversion reaction rate by releasing volatile byproducts, as well as to realize a morphological
control of the perovskite crystallites. Further investigations will be needed to study the
mechanisms behind the formation of mixed halide perovskite crystals.

2.5 Conclusions
In this chapter, a solution-processed sequential deposition method was employed to prepare
perovskite solar cells. First of all, a compact PbI2 film (~140 nm) was prepared by hot casting a
supersaturated PbI2 solution at high spinning speed on a pre-heated substrate. To achieve the
optimum perovskite layer thickness, as well as a stoichiometric reaction condition, different
concentrations of MAI were employed, where the 45 mg/ml MAI solution gave the highest power
conversion efficiency of 7.73%. However, the inferior result in EQE and the large deviation in the
thicknesses between theoretical and experimental data reflect the poor penetration capacity of
MAI into the PbI2 film, which severely limits the conversion into perovskite. Therefore, different
loading times were investigated, but yet no improvement in PCE was gained. A lower MAI
concentration should be used for further loading experiments, which may alleviate the side effect
brought by the excess MA+. The effect of post-annealing time was also investigated, which was
found important to control the crystallinity and surface coverage of the perovskite film. As a
result, the 30-minute annealing gives the best performance. So far, the best device (perovskite
thickness of 232 nm) fabricated from the c-PbI2 shows a PCE of 7.95% and 8.18% in the fast (0.1
V/s) and slow (0.0013 V/s) J-V measurements, respectively. The biased EQE gives a corrected Jsc
of 10.4 mA cm-2 and a PCE of 7.52%.
To solve the blocking and slow conversion issues, porous PbI2 precursor films, were fabricated by
involving a coordinative 4-tert-butylpyridine (TBP) solvent in the first step of sequential
deposition, which enlarged the surface area of PbI2 crystals and facilitated the penetration as well
as reaction rate with MAI solution. The top-view of surface morphology demonstrates the success
of increasing the porosity while all the films remain flat and uniform, by showing that the PbI2
crystals interconnected around the holes without forming large aggregates. The high porosity
nature is also evidenced by the down-shift of the UV-vis absorption spectra while the sharpness
and the onset position (520 nm) of the PbI2 absorption keep at constant. Compared to c-PbI2 films,
porous PbI2 (120 L TBP) displays a much faster conversion rate with MAI. According to the
absorption spectra, the conversion of porous PbI2 is accomplished within 20 s dipping (perovskite
thickness of 262 nm), and the film remains stable when further extends the dipping time to 20
minutes. By optimizing the drying time (for the precursor film) and the post-annealing regime
(for the perovskite layer), the device displays decent PCEs of 7.80% and 8.05% in the fast (0.1
V/s) and slow (0.0013 V/s) J-V measurements, respectively. The bias EQE gives an accurate Jsc
of 13.78 mA cm-2, much higher than that of the c-PbI2 film (10.4 mA cm-2), and result in an
improved PCE of 8.95%. However, compare to the published data,23 the main factor that limits
the PCE is the relatively low current density. Although the perovskite conversion obtained from
p-PbI2 (262 nm) is higher than c-PbI2 film (232 nm), the absorber thickness is still low and
contributes to the decrease of EQE in between 650 nm and 780 nm. To achieve a better
conversion, investigations can be performed on different MAI concentrations and fine-tune the
dipping procedures. Besides, by increasing the concentration as well as lowering the spin speed of
the PbI2xL TBP solution, a thicker porous PbI2 film might be able to produce thicker perovskite
films, however before that, the effect of residual PbI2 on the device performance is to be
investigated. Moreover, it is worth quantifying the porosity of the precursor films obtained from
37

Chapter 2
different TBP additives and drying times, so as to get a better understanding of the conversion
process and to improve the device reproducibility.
To achieve an efficient charge collection process in planar perovskite device, chloride ions are
incorporated to fabricate the mixed halide perovskite film, CH3NH3PbI3-xClx. By using PbCl2 as
the chloride source, mixed precursor solutions of PbCl2 and PbI2 were deposited during the first
step of the sequential deposition, which possibly leads to the formation of a low crystallinity
phase, PbICl, as is evidenced by the left-shifting of the original PbI2 transition in the absorption
spectra with the increasing of PbCl2 concentrations. A small amount of PbCl2 (0.1 M) helps to
increase the surface coverage of mixed precursor film while a porous feature is likely to be
obtained by introducing a larger amount of PbCl2. Very quickly, upon 60 s hot dipping, all the
films display the characteristic perovskite absorption spectra, accompanied by the emergence of
residual PbI2 absorption at 520 nm. This phenomenon relates to a side reaction in which the
iodide exchange with chloride, in the formation of unreacted PbI2. Moreover, when the PbCl2
concentration is larger, though the reaction between PbCl2 and CH3NH3I is accelerated by
releasing the volatile organic byproducts, the roughness of the resultant film also increases. As a
result, the J-V performance of a mixed halide perovskite solar cell is strongly influenced by the
chloride content of the precursor solution. The best device obtained from the 0.1 M PbCl2 + 0.9
M PbI2 solution gives PCEs of 7.97% and 8.04% in the fast (0.1 V/s) and slow (0.0013 V/s) J-V
measurements, respectively. The use of PbCl2 does not induce significant bandgap modification,
and therefore, the main role of chloride is to achieve better control over the conversion reaction
between lead halide and MAI. However, the reaction mechanism is not yet fully understood,
which needs further investigations in combination with other techniques.

2.6 Experiments
Materials and reagents
Unless stated otherwise, all the raw materials and solvents were purchased from commercial
suppliers and used without purification. Methylammonium iodide (CH3NH3I, MAI) was
synthesized by reaction with hydroiodic acid and methylamine, according to a reported procedure.
To avoid the uncertain side effect of humidity, PbI2, PbCl2, MAI, DMF, DMSO, 2-propanol and
4-tert-butylpyridine (TBP) were stored and weighed in an N2 filled glovebox.
Device fabrication
Patterned glass/ITO substrates were cleaned by sonication in acetone for 15 minutes, followed by
hand-washing with soapy water (Dodecyl sulfate sodium salt). Then the substrates were rinsed
with deionized water for 15 min to remove the remnant soap, after which they were sonicated in
the 2-propanol solvent for 15 min and dried with clean, dry N2. Finally, the substrates were
treated in the UV Ozone environment for 30 minutes to remove dust and increase the wettability
of PEDOT:PSS on the substrate. At room temperature (ca. 20 C) and normal laboratory
humidity, a filtered (through a 0.45 filter) poly(3,4-ethylenedioxythiophene):poly
(styrenesulfonic acid) (PEDOT:PSS, Clevios P, VP Al4083) solution was spin-coated at 3000
rpm for 60 s onto a clean substrate and dried on a hot plate at 140 C for 15 minutes. In this
chapter, all the perovskite solar cells (PSCs) were fabricated through the solution processed twostep sequential deposition. Afterwards, an electron-transporting-layer phenyl-C61-butyric acid
methyl ester ([60]PCBM, 24 mg, dissolved in 1 ml 1:1 v/v, chlorobenzene and chloroform) was
spin-coated on top of the perovskite at 500 or1000 rpm for 60 s, with an average thickness of 140
or 100 nm, depending on the high or low roughness of the bottom perovskite layer. Finally, the
substrates were transferred to the thermal evaporator under the pressure of 710 -7 mbar, where
the LiF interlayer of 1 nm and back electrode Al of 100 nm were deposited on top of the devices.
38

Solution-processed sequential deposition for planar perovskite solar cells

The active device area is determined by the overlap of the ITO and Al electrodes, where 0.09 and
0.16 cm2 are used in this study.
Fabrication of methylammonium lead tri-iodide perovskite solar cells
Crystallized PbI2 precursor and MAI deposition: In the wet, N2 filled spin-coating glovebox,
400 mg of PbI2 was dissolved in 1 ml N,N-dimethylformamide (DMF) under stirring overnight at
70 C, and was filtered through a 0.25 m filter before use. The substrates deposited with
PEDOT:PSS layer were transferred to the wet glovebox and pre-heated on the hotplate to 70 C,
after which a hot PbI2 precursor solution (50 l) was placed on top, and the substrates were spincoated at 5700 rpm for 35 s. After spinning, the films were quickly moved back to the hotplate
and dried at 70 C for 20 minutes, to remove all the residual DMF solvent and to produce the
crystallized PbI2 precursor films before cool-down.
In the wet glovebox, MAI with varied concentrations from 25 to 55 mg was dissolved in 1 ml 2propanol at room temperature and filtered before use. During the second deposition step, a 75 l
MAI solution in 2-propanol was placed on top of the cooled PbI2 substrates for several seconds
(loading time, from 0 to 60 s), which was then spin-coated together with the cooled PbI2 film at
5700 rpm for 35 s. The resultant perovskite stacking layers were dried on a hotplate at 100 C for
from 15 to 90 minutes.
Porous PbI2 precursor and MAI dipping: To produce a more porous PbI2 film, x L (x varied
from 0 to 360) TBP was added to 400 mg/ml PbI2 in DMF solution. Due to the stronger
coordinative effect of TBP, PbI2 could dissolve in DMF at room temperature under overnight
stirring. In the wet glovebox, 50 L precursor solution was spin-coated on the PEDOT:PSS
substrate at 5700 rpm for 35 s and followed by drying at 70 C for 30 minutes to evaporate most
of the TBP molecules and to obtain the porous PbI2 morphology, before cooling down to room
temperature.
A large amount of (15 ml) 10 mg/ml MAI in 2-propanol solution was prepared in a beaker. The
PbI2 substrates were fully immersed into the MAI solution for several seconds (specified in
Section 2.3.2), washed with pure 2-propanol and followed by spinning at 4000 rpm for 15 s to dry,
after which the resultant perovskite film was post-annealed at 100 C for several minutes.
Fabrication of methylammonium lead mixed halide perovskite solar cells
Mixed halide precursor film and MAI dipping: Chloride was incorporated by adding x M
PbCl2 into (1-x) M PbI2 precursor dissolved in DMF at 85 C, where x 0.5 due to low solubility
of PbCl2 and the total concentration of the precursor solution was kept at 1.0 M. As a reference,
1.0 M pure PbCl2 in DMSO was also made and filtered before use. Similarly, in the wet glovebox,
a precursor solution (50 l) was spin-coated on top of the pre-heated PEDOT:PSS substrate at
5700 rpm for 35 s and dried at 70 C for 30 minutes, after cooling a mixed halide precursor film
was obtained.
10 mg/ml MAI in 2-propanol solution was prepared in a beaker at 70 C. The mixed halide
precursor substrate was dipped into the MAI solution for 60 s, followed by a rinsing procedure in
pure 2-propanol and spinning at 4000 rpm for 15 s to dry, and finally, a perovskite film was
obtained after post-annealing the substrates at 100 C for 30 minutes.
Device photovoltaic performance characterization
In N2 environment, the J-V curves of the devices were measured under simulated sunlight by
using a tungsten-halogen lamp (~100 mW/cm2, regularly calibrated by a silicon diode) filtered by
a Hoya LB100 daylight filter with a Keithley 2400 source meter. To ensure a more defined device
39

Chapter 2
absorbing area, during the measurements a black mask with two different aperture sizes of 0.0676
and 0.1296 cm2 was placed in front of the substrates, in corresponded to the cell areas of 0.09 and
0.16 cm2, respectively. For the external quantum efficiency (EQE) measurement, simulated AM
1.5G irradiation was generated by using a 50 W tungsten-halogen lamp (Osram 64610) and a bias
green light (530 nm), combining with a mechanical chopper (Stanford Research, SR 540) and a
monochromator (Oriel, Cornerstone 130). The response was recorded as the voltage over a 50
resistance, using a lock-in amplifier (Stanford Research Systems, SR 830). The accurate shortcircuit current density Jsc was determined by integrating the EQE spectra.
The UV-vis absorption spectra were measured by Perkin Elmer UV/Vis/NIR Lambda 900
spectrometer. Samples were deposited on top of glass substrates (without ITO electrode),
followed by the same deposition processes as solar cells.
Film characterization
A Veeco Dektak150 profilometer measured the film thickness and roughness. Surface
morphology of all kinds of precursor and perovskite films were recorded using a Focused-ionbeam scanning electron microscope (FIB-SEM, 5 kV, FEI company).

40

Solution-processed sequential deposition for planar perovskite solar cells

2.7 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.

H. Kim, K.-G. Lim and T.-W. Lee, Energy Environ. Sci., 2016, 9, 12-30.
L. Zheng, D. Zhang, Y. Ma, Z. Lu, Z. Chen, S. Wang, L. Xiao and Q. Gong, Dalton
Trans., 2015, 44, 10582-10593.
S. D. Stranks, P. K. Nayak, W. Zhang, T. Stergiopoulos and H. J. Snaith, Angew. Chem.
Int. Ed., 2015, 54, 3240-3248.
H. Zhou, Q. Chen and Y. Yang, MRS Bulletin, 2015, 40, 667-673.
J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M.
Gratzel, Nature, 2013, 499, 316-319.
D. Liu and T. L. Kelly, Nat. Photonics, 2013, 8, 133-138.
Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang, Y. Gao and J. Huang,
Energy Environ. Sci., 2014, 7, 2619.
J. Seo, S. Park, Y. Chan Kim, N. J. Jeon, J. H. Noh, S. C. Yoon and S. I. Seok, Energy
Environ. Sci., 2014, 7, 2642.
Q. Lin, A. Armin, R. C. R. Nagiri, P. L. Burn and P. Meredith, Nat. Photonics, 2014, 9,
106-112.
D. Liu, M. K. Gangishetty and T. L. Kelly, J. Mater. Chem. A, 2014, 2, 19873-19881.
S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. Alcocer, T. Leijtens, L. M.
Herz, A. Petrozza and H. J. Snaith, Science, 2013, 342, 341-344.
T.-B. Song, Q. Chen, H. Zhou, C. Jiang, H.-H. Wang, Y. Yang, Y. Liu, J. You and Y.
Yang, J. Mater. Chem. A, 2015, 3, 9032-9050.
T. Zhang, M. Yang, Y. Zhao and K. Zhu, Nano Lett., 2015, 15, 3959-3963.
C.-G. Wu, C.-H. Chiang, Z.-L. Tseng, M. K. Nazeeruddin, A. Hagfeldt and M. Grtzel,
Energy Environ. Sci., 2015, 8, 2725-2733.
H. J. Snaith, J. Phys. Chem. Lett., 2013, 4, 3623-3630.
Z. Tan, W. Zhang, D. Qian, C. Cui, Q. Xu, L. Li, S. Li and Y. Li, Phys. Chem. Chem.
Phys., 2012, 14, 14217-14223.
H. A. Harms, N. Tetreault, N. Pellet, M. Bensimon and M. Gratzel, Faraday Discuss.,
2014, 176, 251-269.
Y. Cheng, H.-W. Li, J. Zhang, Q.-D. Yang, T. Liu, Z. Guan, J. Qing, C.-S. Lee and S.-W.
Tsang, J. Mater. Chem. A, 2015, 4, 561-567.
A. Dualeh, N. Ttreault, T. Moehl, P. Gao, M. K. Nazeeruddin and M. Grtzel, Adv.
Funct. Mater., 2014, 24, 3250-3258.
G. E. Eperon, V. M. Burlakov, P. Docampo, A. Goriely and H. J. Snaith, Adv. Funct.
Mater., 2014, 24, 151-157.
G. Li, T. Zhang and Y. Zhao, J. Mater. Chem. A, 2015, 3, 19674-19678.
H. Zhang, J. Mao, H. He, D. Zhang, H. L. Zhu, F. Xie, K. S. Wong, M. Grtzel and W. C.
H. Choy, Adv. Energy Mater., 2015, 5, 1501354-1501364.
Y. Shi, X. Wang, H. Zhang, B. Li, H. Lu, T. Ma and C. Hao, J. Mater. Chem. A, 2015, 3,
22191-22198.
Y. Wu, A. Islam, X. Yang, C. Qin, J. Liu, K. Zhang, W. Peng and L. Han, Energy
Environ. Sci., 2014, 7, 2934.
Y. Xie, F. Shao, Y. Wang, T. Xu, D. Wang and F. Huang, ACS Appl. Mater. Interfaces,
2015, 7, 12937-12942.
M. I. Dar, M. Abdi-Jalebi, N. Arora, M. Grtzel and M. K. Nazeeruddin, Adv. Energy
Mater., 2015, 1501358.
Y. Ma, L. Zheng, Y. H. Chung, S. Chu, L. Xiao, Z. Chen, S. Wang, B. Qu, Q. Gong, Z.
Wu and X. Hou, Chem. Commun., 2014, 50, 12458-12461.
S. Dharani, H. A. Dewi, R. R. Prabhakar, T. Baikie, C. Shi, D. Yonghua, N. Mathews, P.
P. Boix and S. G. Mhaisalkar, Nanoscale, 2014, 6, 13854-13860.
41

Chapter 2
29.
30.
31.
32.
33.
34.
35.

M. Wang, C. Shi, J. Zhang, N. Wu and C. Ying, Journal of Solid State Chemistry, 2015,
231, 20-24.
Y. Tidhar, E. Edri, H. Weissman, D. Zohar, G. Hodes, D. Cahen, B. Rybtchinski and S.
Kirmayer, J. Am. Chem. Soc., 2014, 136, 13249-13256.
Y. Li, W. Sun, W. Yan, S. Ye, H. Peng, Z. Liu, Z. Bian and C. Huang, Adv. Funct.
Mater., 2015, 25, 4867-4873.
N. Wu, C. Shi, C. Ying, J. Zhang and M. Wang, Appl. Surf. Sci., 2015, 357, 2372-2377.
H. Yu, F. Wang, F. Xie, W. Li, J. Chen and N. Zhao, Adv. Funct. Mater., 2014, 18721879.
C. Liu, J. Fan, X. Zhang, Y. Shen, L. Yang and Y. Mai, ACS Appl. Mater. Interfaces,
2015, 7, 9066-9071.
H. J. Snaith, A. Abate, J. M. Ball, G. E. Eperon, T. Leijtens, N. K. Noel, S. D. Stranks, J.
T. Wang, K. Wojciechowski and W. Zhang, J. Phys. Chem. Lett., 2014, 5, 1511-1515.

42

Chapter 3 Vapor-assisted solution process for planar perovskite solar


cells

Abstract
A new method named vapor-assisted solution process was investigated to fabricate perovskite
solar cells, which uses a thermal gradient sublimer to sublimate the CH3NH3I powder to the gas
phase in a dedicated zone, and followed by reacting with the solution-processed lead halide
substrates, based on a low-pressure condition. Side effects arose from the solution deposition can
be circumvented and therefore, resulting in a much-enhanced controllability of perovskite film
quality. Both the compact and porous PbI2 precursor films were employed for the vapor
deposition methods, and the competition between chemical conversion reactions and physical
vapor deposition processes was systematically investigated. Under 100 mW cm-2 illumination
condition, the best-performing device fabricated from the compact PbI2 substrate displays a
power conversion efficiency of 7.65%, while the corrected PCE of a porous PbI2 fabricated
device is 7.04%.

Chapter 3

3.1 Introduction
Although possessing the advantages of inexpensive, low temperature and easy processability, a
solution-based sequential deposition method is mainly successful in fabricating the mesoporous
perovskite solar cells.1 Due to the lack of sufficient surface area, the exposure of the lead halide
precursor film to organic ammonium halide solution is severely limited, the thickness of the flat
perovskite layer is often limited to 300 nm.2 As a result, the power conversion efficiency (PCE)
of a planar perovskite cell fabricated by a solution-sequential deposition process is yet
unsatisfying.3
As is discussed in Chapter 2, to grow a thick, planar perovskite film, a long reaction time is
desired to reach a complete conversion of the lead halide. However, in a solution-processed
sequential deposition, the dissolution of perovskite films in the solution phase leads to an
increased surface roughness and reduced PV performance. Therefore, a vapor-assisted solution
process (VASP) was developed,4 which involves the in-situ reaction between the volatile organic
vapor and the solid pre-deposited lead halide film. Combining the advantages of the template
effect of lead halide precursor as well as the absence of solvent media, in the first VASP paper, 4
the device already exhibits an outstanding PCE of 12.1%.
In this section, to explore the feasibility of this new method, a thermal gradient sublimer was used
for generating CH3NH3I vapors while the substrates pre-deposited with PbI2 were placed at a
target area for the conversion to perovskite.

3.2 VASP using a thermal gradient sublimer

Figure 3.1 A schematic of the evaporation process in a thermal gradient sublimer.

A thermal gradient sublimer is usually applied for the purification of small molecular weight
materials and the formation of large crystals.5, 6 Fig. 3.1 shows a schematic diagram of the vapor
deposition process. Firstly, a lead halide precursor film is deposited onto the substrate. Then the
substrates are placed in a sequence at the cooling part of the tube while the CH3NH3I material is
put in the sample boat. After sealing, the quartz tube keeps pumping down (ca. 0.50 mbar)
through the open end and is placed in the furnace of the sublimer. By heating up to a certain
temperature, CH3NH3I evaporates and travels down to the cooling region, where the pressure
exceeds its vapor pressure and materials start to deposit on both of the substrates and the cold
wall of the tube. CH3NH3I molecules start to intercalate into the PbI2 layer and result in the
formation of perovskite crystals, as is evidenced by the gradual color change of the substrate from
bright yellow to dark brown. However, it is found that at a distance close to the sample boat, the
deposition rate of CH3NH3I vapor is already fast, and an excess amount of CH3NH3I powders is
seen to deposit both on the wall of the quartz tube as well as the substrates. This results in a
much-increased film roughness, see the projections in SEM images of Fig. 3.2. Besides, as the
concentration of organic compound gradually decreases towards the open end, the deposition rate
44

Vapor-assisted solution deposition for planar perovskite solar cells

significantly decreases downward the tube. Therefore, the substrates placed far from the sample
boat take much more time to obtain the dark-brown color of perovskite films.

Figure 3.2 Top-view SEM images of perovskite films with excess CH3NH3I deposited on top, fabricated via vaporassisted solution process at 180 C (sample boat) for 20 minutes.

Three steps control the formation of perovskite film in VASP: Evaporation and transportation of
CH3NH3I material in the gas phase, physical deposition of CH3NH3I in the gas/solid phase, the
chemical reaction of CH3NH3I with PbI2. These steps rely on the precise control of pressure and
temperature. During the deposition process, the quartz tube was kept on pumping at a high
vacuum degree to prevent the contamination from the ambient environment. The evaporation rate
of CH3NH3I dependents on the chamber pressure and the temperature of sample boat, according
to Equation 3.1:7, 8

5.84 102 (

M 1/2
) ( Pe Ph )
T

g cm 2 s 1

(3.1)

Where is the mass evaporation rate of CH3NH3I, M is the molecular weight, T is the
temperature of the sample boat, Pe is the vapor pressure of CH3NH3I at T, and Ph is the pressure
inside the tube. Under a particular Ph, the evaporation rate is mainly influenced by the
temperature T. At a higher source temperature T, the vapor pressure of CH3NH3I increases and
higher evaporation rate of CH3NH3I can be achieved, accordingly the concentration of CH3NH3I
and the diffusion rate of the gas phase also increases. However, since the sublimer supplies only
one heater at the sample boat, the rest part of the tube is cooled down, a higher T of the sample
boat only slightly increases the temperature of deposition region. The diffusion rate of a dilute gas
phase is given by 7, 9

Dg T 3/2 / P

(3.2)

Where Dg is the gas phase diffusion rate. As a result, the vapor CH3NH3I molecules can only
diffuse to a short distance from the sample boat, and deposit fiercely due to the low temperature
of the tube, while it takes much longer time for substrates far from the sample boat to deposit the
requisite amount of CH3NH3I. Although a cold PbI2 substrate favors the physical deposition of
organic vapor molecules, the low temperature depresses the gas/solid phase diffusion process
which relates to the penetration of the organic molecules into the PbI2 film, according to the
diffusion equation:9

45

Chapter 3

Ds Ds0 exp(

EaSurf
), Eas
kT

Eabulk

(3.3)

Where Ds is the diffusion rate of the solid CH3NH3I. Similarly, consider the chemical reaction
kinetics, the reaction rate for the perovskite formation process is given by10

v k [CH 3 NH 3I][PbI 2 ]

(3.4)

k A exp Ea / RT

(3.5)

Moreover, the rate constant k is:

The formation of hybrid perovskite crystals is an endothermal reaction, with Ea > 0.10 Therefore, a
relatively high substrate temperature is desired for an efficient solid diffusion process and
chemical reaction between CH3NH3I and the PbI2, and to avoid the over-rapid physical vapor
deposition. In the meantime, a high temperature of the sample boat and a low chamber pressure
accelerate the vapor evaporation, gas phase diffusion, and chemical reaction rate.

3.2.1 Vapor deposition on c-PbI2 (compact) substrates


Follow the same procedure in Section 2.2; a compact PbI2 layer was deposited onto the
Glass/ITO/PEDOT:PSS substrate. Then these PbI2 substrates were loaded alongside each other in
the deposition region of the quartz tube, and CH3NH3I powders amounted from N2 glove box
were loaded into the sample boat. In the preliminary experiments, after pumping down to 0.3
mbar, the sample boat was heated to a nominal 180 C, while the stabilized temperature on a
substrate (the one placed close to the sample boat) was only ~110 C, measured by a thermometer.
This substrate temperature is lower than the minimum limit (ca. 140 C, combine with a chamber
pressure of 0.01 mbar) for the fabrication of efficient perovskite solar cells, suggested by the
others9, 11 who use the double-zone furnace for VASP method. As is discussed in Chapter 2,
annealing temperatures higher than 90 C are sufficient to fulfill the energy needs of the chemical
reaction of perovskite compounds. Therefore, when the CH3NH3I material is set at 180 C, the
relatively high evaporation rate brings a considerable amount of CH3NH3I molecules to the PbI2
surface (the substrates close to the sample boat) and induces a quick reaction between the two
components. This is evidenced by the rapid change in color of the substrates, from yellow to dark
brown, within 5 minutes. On the other hand, under the chamber pressure of 0.3 mbar and a
substrate temperature of 110 C, it is also thermodynamically favorable for vapor phase of
CH3NH3I to reach the PbI2 substrate (compared to the substrate temperature of 140 C and 0.01
mbar scenario9) and gives rise to excessive physical deposition processes. It is observed that, after
10 minutes deposition, excess white CH3NH3I powders start to deposit onto the dark brown film,
which dramatically increases the surface roughness. As a result, under 180 C and 160 C
evaporation conditions, the deposition processes were hard to control, and no working devices
have been fabricated.
Since it was not feasible to increase the substrate temperature, the temperature of the sample boat
with CH3NH3I was lowered to 140 C, which decreased the evaporation rate and less CH3NH3I
vapor molecules would present at the PbI2 surface, in the hope of achieving a milder reaction
condition for the perovskite film. However in the meantime, the substrate temperature also
dropped to 90 C, measured by a thermometer. On this occasion, a sequence of substrates covered
with 160 nm thick PbI2 films was evaporated with CH3NH3I at 140 C for 45 minutes in the
46

Vapor-assisted solution deposition for planar perovskite solar cells

sublimer (under 0.35 mbar). The resultant thickness of the third substrate is 269 nm,
corresponding to the best device (without post-annealing) efficiency at the 3a2 cell. As is shown
in Fig. 3.3(a), under 100 mW cm-2 light illumination and the fast J-V measurements (0.01 V/s),
the 3a2 device displays a PCE of 7.65% in the downward scanning direction, while a PCE of
6.9% is obtained in the upward scanning direction. In the slow J-V measurement (0.0013 V/s), the
device displays a poorer short-circuit current density which leads to a lower PCE of 6.42%. Fig.
3.3(b) shows the current density and PCE of the 3a2 cell as a function of time at Vmpp (0.79 V),
the stabilized power output is close to 7%.
(b) 10

(a)

Downward, 7.65 %
Upward, 6.9%
Slow sweep, 6.42%

(%), Jmax (mA cm-2)

Current density J (mA/cm2)

-4

-8

PCE(%)
Jmax(mA cm-2)

5
4
3
2
1
0

-12
0.0

0.2

0.4

0.6

0.8

1.0

Voltage (V)

100

200

300

400

500

600

Time (s)

Figure 3.3 PV performances of the 3a2 cell, under 100 mW cm-2. (a) J-V measurements at fast sweep (0.1 V/s), ()
open-circuit to short-circuit direction and () reverse direction; () measured at slow sweep (0.0013 V/s). (b) Jmpp and
PCE as a function of time, the applied voltage is close to the Vmpp, 0.79 V.
Table 3.1 Photovoltaic parameters of devices deposited at 140 C for 45 minutes in a thermal gradient sublimer.
Substrate No. 1 to 4 is placed in a sequence at the deposition region, from the left to the right of the quartz tube.

Substrate
No.
1

Cell

Jsc (mA cm-2)

Voc (V)

FF

PCE (%)

a1
b1
a2
b2
a1
b1
a2
b2
a1
b1
a2
b2
a1
b1
a2
b2

2.03
6.43
1.27
1.59
3.69
7.78
2.59
2.32
7.24
10.20
10.70
7.00
1.72
1.28
1.46
1.21

0.924
0.780
0.885
0.937
0.947
0.931
0.977
0.943
0.939
0.976
0.975
0.933
0.727
0.749
0.767
0.692

0.585
0.529
0.668
0.646
0.608
0.553
0.594
0.671
0.392
0.690
0.731
0.390
0.389
0.344
0.418
0.324

1.10
2.65
0.75
0.96
2.13
4.00
1.51
1.47
2.66
6.90
7.65
2.55
0.49
0.33
0.47
0.27

The efficiencies of the whole batch of cells are given in Table 3.1. It is remarkable that the device
efficiencies are sparsely distributed, which is an indication of the uneven deposition and reaction
processes downward the tube. It was observed that after 45 minutes of evaporation, a thin layer of
white floccule CH3NH3I was deposited on the surface of the first two substrates. At each
47

Chapter 3

evaporation temperature, to ensure the highest conversion of PbI2 into perovskite, the vapor
deposition process continues up until the first substrate gets extra CH3NH3I powders deposited
while the second/third substrate remains the mirror-like, dark brown appearance. Before further
processing, all the substrates were washed with isopropanol to remove the residual CH3NH3I.

Figure 3.4 Top-view SEM images of perovskite film fabricated from VASP, at (a) 120 C for 100 minutes, (b) 130 C
for 100 minutes, and 140 C for 45 min (c) without post-annealing, (d) with post-annealing at 100 C for 15 min.

Furthermore, the effect of evaporation temperatures on the final device performance was
investigated. The vapor pressure decreases with evaporation temperature. As a result, fewer
CH3NH3I molecules will be present on top of the PbI2 substrate, and the reaction between PbI2
and CH3NH3I becomes less drastic. However, the temperature of the substrate also decreases with
the source temperature, which makes it difficult to control the balance between the physical vapor
deposition and the chemical reaction. Therefore, it is not feasible to achieve a full conversion of
PbI2 using this method. Fig. 3.4 shows the top-view SEM images of perovskite films prepared by
evaporating at (a) 120 C for 100 minutes, (b) 130 C for 100 minutes and 140 C for 45 min (c)
without thermal annealing or (d) with further annealing at 100 C for 15 minutes, respectively.
Compare to (a) and (b), 140 C evaporation produce the finest surface morphology, which might
be attributed to the improved conversion of PbI2. After post-annealing of perovskite film at
100 C for 15 minutes, the grain sizes and surface coverage slightly increases.
Fig. 3.5 shows the J-V characteristics of the perovskite solar cells fabricated by evaporating at
120 C, 130 C, and 140 C, as summarized in Table 3.2. The resultant PCE shows a random
trend, which decreases from 4.12% (Jsc = 7.29 mA cm-2, Voc = 0.980 V) to 3.03% (Jsc = 5.42 mA
cm-2, Voc = 0.945 V) when evaporation temperature increases from 120 C to 130 C, and then
increases to 7.65% (Jsc = 10.70 mA cm-2, Voc = 0.9750 V) at 140 C evaporation. The difference
48

Vapor-assisted solution deposition for planar perovskite solar cells

in PCE is mainly derived from different Jsc, due to the uncontrolled conversion of PbI2 to
perovskite at each deposition condition. Moreover, although displaying a good surface
morphology, the post-annealing treatment for perovskite film results in the decreases of Jsc and
Voc, and a drop of PCE from 7.65% to 3.83%. The fluctuating variation for the PCEs reflects the
poor reproducibility of the vapor deposition method via the thermal gradient sublimer.
120 C
130 C
140 C
140 C, anneal at 100 C for 15 min.

Current density J (mA/cm2)

-4

-8

-12
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

Voltage (V)

Figure 3.5 J-V curves of the perovskite devices fabricated at different evaporation temperatures.
Table 3.2 Photovoltaic performances of best-performing devices with different evaporation conditions. Measured at
100 mW cm-2 illumination condition.

Evaporation T
120 C
130 C
140 C
140 C, annealed

Jsc (mA cm-2)


7.29
5.42
10.70
7.50

Voc (V)
0.980
0.945
0.975
0.965

FF
0.577
0.592
0.731
0.529

PCE (%)
4.12
3.03
7.65
3.83

3.2.2 Vapor deposition on p-PbI2 (porous) substrates


Drawn from Section 3.2.1, it is concluded that the physical vapor deposition rate is higher than
the chemical reaction rate. At the very start of the evaporation, newly deposited CH 3NH3I can
equally react with PbI2 on the surface and leads to the formation of a compact perovskite film
(see Fig. 3.4). As the experiments proceeds, however, extra CH3NH3I starts to deposit onto the
film which results in the formation of a rough and insulating floccule layer. Due to the relatively
low temperature of the bottom substrate, diffusion of the organic solid phase is severely hindered
and therefore, the conversion of PbI2 to perovskite is limited. Therefore, it seems impossible to
fabricate efficient perovskite cells via a thermal gradient sublimer, by simply increasing the
reaction time between CH3NH3I vapor and the PbI2 substrate. Similar to the discussion in Chapter
2, one promising solution is to increase the porosity of the PbI2 precursor film, by which the
surface area gets enlarged to achieve a better organic vapor diffusion as well as efficient reactions
between CH3NH3I and PbI2. During the first step deposition, a precursor solution of 400 mg/ml
PbI2120 L TBP in DMF was spin coated onto a Glass/ITO/PEDOT:PSS substrate to produce
the porous PbI2 morphology (see Fig. 2.9). Then the substrates were aligned in the sublimer and
started the deposition.

49

Chapter 3

Table 3.3 Photovoltaic parameters of devices deposited at 140 C for 60 minutes. Substrate No. 1 to 4 was placed in a
sequence at the deposition region, from the left to the right of the quartz tube.

Substrate
No.
1

Cell

Jsc (mA cm-2)

Voc (V)

FF

PCE (%)

a1
b1
a2
b2
a1
b1
a2
b2
a1
b1
a2
b2
a1
b1
a2
b2

1.88
2.71
1.74
2.87
9.31
7.47
6.76
9.48
10.70
9.92
10.00
10.10
3.93
4.08
3.65
3.01

0.860
0.854
0.751
0.894
0.893
0.864
0.844
0.890
0.714
0.699
0.739
0.662
0.312
0.357
0.341
0.299

0.672
0.724
0.413
0.652
0.777
0.734
0.703
0.764
0.477
0.362
0.329
0.403
0.289
0.296
0.296
0.295

1.09
1.68
0.54
1.67
6.46
4.74
4.01
6.45
3.66
2.51
2.44
2.69
0.36
0.43
0.37
0.27

(a) 4

(b) 0.8
2

No bias, 12.17 mA/cm


-2
Biased, 10.69 mA/cm

0.7

0.6
0.5

-4

EQE

Current density J (mA cm-2)

Slow sweep, 6.1%


Fast sweep, 6.46%

0.4
0.3

-8

0.2
0.1

-12
0.0

0.2

0.4

0.6

0.8

0.0

1.0

300

Voltage (V)

400

500

600

700

800

Wavelength (nm)

Figure 3.6 J-V curves and EQE for the best-performing 2a1 cell. (a) Under 100 mW cm-2 illumination, measurements
were taken at fast (0.1 V/s) and slow (0.0013 V/s) sweep. (b) EQE spectra, () under bias light (530 nm green light)
and () without bias light.

Similar to Section 3.2.1, the best device is produced by evaporating CH3NH3I at 140 C for 60
minutes with a set of porous, 146 nm thick PbI2 substrates (ca. 90 C), without post-annealing.
Table 3.3 shows the efficiencies of the whole batch of cells, where the 2a1 cell gives the highest
efficiency with a perovskite layer thickness of only 210 nm, demonstrating that the porous PbI2
do not aid the conversion of the substrates. Fig. 3.6 displays the J-V curve and EQE spectra of the
2a1 cell. Under 100 mW cm-2 light illumination, the device shows a PCE of 6.46% in the fast J-V
measurement (0.1 V s-1), with Jsc = 9.31 mA cm-2, Voc = 0.893 V and FF = 0.777. The slow J-V
measurement (0.0013 V s-1) gives an analogous PCE of 6.10%, yielding a Jsc of 9.24 mA cm-2, Voc
of 0.860 V and FF of 0.766. The hysteresis between both J-V measurements is small. The
integrated Jsc of 10.69 mA cm-2 from the biased EQE spectrum (530 nm bias light, AM 1.5G
illumination) is slightly higher than the J-V measured value, which gives a corrected PCE of
7.04%. It is notable that, compare to the wide distribution in PCEs shown in Table 3.2, the PCEs
of devices fabricated from p-PbI2 substrate only show a small deviation. Also, it is found that the
50

Vapor-assisted solution deposition for planar perovskite solar cells

deposition of an insulating CH3NH3I layer on the perovskite film severely reduces the Jsc of the
devices, as is evidenced by Substrate 1 shown in Table 3.3. In the meantime, it is observed that
the color of Substrate 4 after evaporation for 60 minutes is slightly brown, which means that the
conversion of PbI2 is tiny since most of the CH3NH3I molecules get deposited on Substrate 1 (and
2). According to Table 3.3, it is found that a lower conversion of PbI2 results in a much decreased
Jsc and Voc.
The effect of the p-PbI2 layer thickness, as well as the post-annealing on the device performance,
are summarized in Fig. 3.7 and Table 3.4. After evaporation at 140 C for 60 minutes, the device
fabricated by a thicker p-PbI2 precursor film (194 nm) displays a much-lowered PCE and Jsc, with
a perovskite layer thickness of only 230 nm, indicating that it is the incomplete conversion of
PbI2 that limits the device performance. Moreover, the device will present more insulating PbI2
near the PEDOT:PSS layer, which hinders the charge collection and contributes to a lower Jsc.12
Also, compared to the device without thermal annealing, after annealing at 100 C for 30 minutes,
the Jsc increases to 11.63 mA cm-2 while Voc decreases to 0.671 V. The increase in Jsc could be
attributed to the better crystallinity of the perovskite film, and the reduce in Voc is due to the
formation of large sizes of pinholes. Fig. 3.7(b) illustrates a considerable bias dependence in the
EQE spectra of the post-annealed device. Under a bias light illumination (with the 530 nm green
light source), the inferior perovskite morphology will induce enormous charge recombination,
leading to a lower incident photon-to-current efficiency.
140 C, 60 min., wo anneal
140 C, 60 min., anneal at 100 C for 30 min.
140 C, 60 min. (thick precursor: 194 nm)
120 C, 110 min., wo anneal
120 C, 180 min., wo anneal

(b) 0.8
No bias, 10.85 mA/cm
2
Biased, 1.83 mA/cm

0.6
0.5

-4

0.4
0.3

-8

0.2
0.1

-12

-0.2

0.0

0.2

0.4

0.6

0.8

0.0
300

1.0

Voltage (V)
(c) 0.8
0.7

No bias, 10.93 mA/cm


2
Biased, 6.40 mA/cm

0.7

0.6

0.6

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0
300

400

500
600
Wavelength (nm)

700

800

(d) 0.8
Evaporation: 120 C, 110 min.
No post-annealing

EQE

EQE

Evaporation: 140 C, 60 min.


Post-annealing: 100 C, 30 min.

0.7

EQE

Current density J (mA/cm2)

(a)

400

500
600
Wavelength (nm)

700

0.0
300

800

Evaporation: 120 C, 180 min.


No post-annealing

400

500
600
Wavelength (nm)

No bias, 11.66 mA/cm


2
Biased, 9.29 mA/cm

700

800

Figure 3.7 Effects of precursor thickness, evaporation temperature/time and post-annealing of the perovskite film on
the final device performance. (a) J-V curves of the highest-performing devices, under 100 mW cm-2 illumination. EQE
spectra of (b) 100 C for 30 minutes post-annealed perovskite device, evaporated at 140 C for 60 minutes, (c)
evaporated at 120 C for 110 minutes, without post-annealing, and (d) evaporated at 120 C for 180 minutes, without
post-annealing.

51

Chapter 3

Table 3.4 Photovoltaic performances of devices with different evaporation conditions, determined at 100 mW cm-2
illumination condition. All the PV parameters are averaged values from four cells on one substrate.

Evaporation T
120 C, 110 min.
120 C, 180 min.
140 C, 60 min.
140 C, annealed
140 C, thicker p-PbI2

Jsc (mA cm-2)


9.37
8.35
8.26
11.63
2.22

Thickness (nm)
186 nm
202 nm
210 nm
207 nm
230 nm

Voc (V)
0.735
0.885
0.873
0.671
0.862

FF
0.489
0.643
0.745
0.508
0.659

PCE (%)
3.36
4.75
5.42
3.99
1.25

When the evaporation temperature of CH3NH3I decreases to 120 C, after 110 minutes, the
devices give a slightly higher average Jsc of 9.37 mA cm-2, but a much decreased Voc, FF and PCE,
which are 0.735 V, 0.489 and 3.36%, respectively. When prolonging to 180 minutes, a Jsc of 8.35
mA cm-2, Voc of 0.885 V, FF of 0.643 and PCE of 4.75% are achieved in average, respectively.
However, the surface of Substrate 1 in that batch is covered with white floccule material, which
indicates that the reaction between CH3NH3I and the PbI2 cannot proceed anymore. In Fig. 3.8, (a)
and (b) display the top-view SEM images of the vapor deposited perovskite film for 110 and 180
minutes, respectively. It is found that a longer deposition time leads to the formation of larger
grain sizes accompanied by the fully covered surface of the perovskite film, which might improve
the charge transporting ability and contribute to a higher Voc and FF. As is depicted in Fig. 3.7(c)(d), the bias dependence of the 180 minutes deposited device is much smaller than the 110
minutes deposited device. It is also an indication of the reduced charge recombination, due to the
existence of fewer defects in a long-time deposited perovskite film.
In summary, devices fabricated from the porous precursor films display an even distribution of
PCEs on each substrate, and result in a better reproducibility from batch to batch via the vapor
deposition methods. However, the Voc of the solar cells is smaller than that of devices fabricated
from the c-PbI2 substrates, which still could be attributed to the incomplete conversion into
perovskite. Therefore, the perovskite layer is less compact and deteriorates the final PV
performances. It is worthwhile to further optimize the porosity of the p-PbI2 films, to reach a
more evenly distributed PCE as well as a compact perovskite morphology.

Figure 3.8 SEM images of perovskite film evaporated at 120 C for (a) 110 min., (b) 180 min., respectively.

52

Vapor-assisted solution deposition for planar perovskite solar cells

3.3 VASP with home-build equipment


In the VASP, the substrate temperature is the most critical parameter that determines the growth
of the perovskite film, by controlling the balance between physical deposition and chemical
conversion reaction. As is shown in Section 3.2, in a thermal gradient sublimer, the substrate
temperature is much lower compared to the temperature of the sample boat, which leads to a high
sticking coefficient of CH3NH3I and physical vapor deposition rate.13 Meanwhile, the chemical
reaction of the two components is relatively slow at a low substrate temperature, therefore
resulting in an incomplete conversion of perovskite with insulating contaminants on top of the
devices, reducing the PV performances. Recently, a home-designed two-zone furnace was built
for a precise temperature control during the evaporation experiments. The appearance of device
setup and the temperature-time graphs can be found in Appendix 1. The new equipment can give
better control over all temperature parameters, as well as the atmosphere involved in the
perovskite film formation. The solid CH3NH3I is placed in the crucible before heating to a
nominal high-temperature to convert to the vapor phase, in the meantime, the whole furnace pipe
covered with the thermal insulating material is heating to the same temperature, to avoid the
deposition of CH3NH3I on the cold wall. The sample holder keeps at the desired deposition
temperature while the organic vapor travels downstream in the tube and starts to react with the
substrates.

3.4 Conclusions
In this chapter, a vapor-assisted solution process was investigated to prepare the perovskite solar
cells. By using a similar procedure in Chapter 2, a compact p-PbI2 precursor film was firstly
deposited onto a Glass/ITO/PEDOT:PSS substrate via the solution process. The following vapor
deposition processes were conducted via a thermal gradient sublimer, where the p-PbI2 substrates
were placed in a sequence in the cooling part of the long quartz tube while the volatile organic
CH3NH3I material is put in the sample boat and heated after pumping down. The best device was
prepared by evaporating CH3NH3I at 140 C for 45 minutes onto the 160 nm thick c-PbI2
substrates, which gives a PCE of 7.65% in the fast downward J-V measurement, and a stabilized
power output close to 7% after 600 seconds operating at maximum power point. However, the
efficiencies of the whole batch of substrates are sparsely distributed, demonstrating the uneven
deposition and reaction processes in the thermal gradient sublimer. Usually, the first substrate
gets the most rapid vapor deposition and a rough surface, while substrates after the third need
very long time to get the conversion to dark-brown color. The effect of lower evaporation
temperatures (120 C, 130 C) on the final device performances was also investigated. A random
variation trend in the efficiencies is observed for different batches of cells, which might be
attributed to the uncontrolled conversion of PbI2. By decreasing the evaporation temperature, the
vapor pressure decreases, and the reaction between PbI2 and CH3NH3I becomes less drastic,
however in the meantime, the substrate temperature also drops, which favors the physical
deposition instead of chemical reactions.
Since it is not possible to increase the conversion degree by simply increasing the reaction time
between CH3NH3I vapor and the PbI2 substrate, porous p-PbI2 substrates were fabricated to
accelerate the vapor diffusion process. Similarly, the best device was obtained by evaporating
CH3NH3I at 140 C for 60 minutes onto the 146 nm thick p-PbI2 substrates, which displays a PCE
of 6.46% in the fast (0.1 V s-1) J-V measurement, and a corrected PCE of 7.04% from the biased
EQE spectrum. Compared to the broad distribution in PCEs on each c-PbI2 substrate, the
53

Chapter 3

efficiencies on each p-PbI2 substrate only show a small deviation, indicating that the porous
precursor film has equal probability to react with CH3NH3I and results in better reproducibility.
However, since the conversion is still limited, the Voc is reduced due to a less compact perovskite
film and therefore, the PCE is also lower. After post-annealing, the device displays a considerable
bias dependence in biased EQE, indicating an inferior perovskite morphology.

3.5 Experiments
Unless stated otherwise, all the raw materials and solvents were used as received.
Methylammonium iodide was synthesized via a reported procedure.
The substrate cleaning procedure, deposition of the hole transporter PEDOT:PSS and electron
transporter [60]PCBM, as well as the metal evaporation of back Al electrode, are followed by the
same method as stated in Section 2.6.
C-PbI2 precursor films:
As is mentioned in Section 2.6, a 400 mg/ml pure PbI2 in DMF solution was prepared in an N2
filled glovebox, and hot cast at 4000 rpm for 35 s onto the PEDOT:PSS covered Glass/ITO
substrates, followed by a drying process at 70 C for 20 minutes to produce a compact PbI2
precursor film.
P-PbI2 precursor films:
120 L 4-tert-butylpyridine (TBP) was added to prepare a 400 mg/ml PbI2 in DMF solution.
After spin-coating the solution onto a Glass/ITO/PEDOT:PSS substrate, and dried at 70 C for 30
minutes, a porous PbI2 precursor film is obtained.
Vapor deposition of methylammonium iodide
Substrates pre-deposited with either a compact or porous PbI2 layer were placed in a sequence,
from the far left of the deposition region of the long quartz tube to the right. The CH 3NH3I white
powder was loaded into the sample boat and covered by glass fiber, in the case of splashing
during pumping down. Then the glass tube was sealed and inserted into the furnace of a thermal
gradient sublimer. The open end of the tube was connected to a vacuum pump, and pumped down
to a pressure of ~0.3 mbar for 15 minutes, before starting to heat the CH3NH3I material. The
sample boat was ramped up to a temperature of 100 C, to warm-up the deposition region where
the substrates were located, after which the temperature was increased to 120 C ~ 140 C for a
different period until the substrates displayed a dark-brown color. The substrates were then
removed from the furnace, washed with pure 2-propanol, and spin at 4000 rpm for 15 s to dry.
The resultant perovskite films were either annealed on a hotplate at 100 C for several minutes or
without undergoing the post-annealing.
Device photovoltaic performance characterization
The J-V characteristics and EQE of perovskite solar cells are measured by the same set-up as
stated in Section 2.6.

54

Vapor-assisted solution deposition for planar perovskite solar cells

3.6 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

J. Burschka, N. Pellet, S. J. Moon, R. Humphry-Baker, P. Gao, M. K. Nazeeruddin and M.


Gratzel, Nature, 2013, 499, 316-319.
Z. Xiao, C. Bi, Y. Shao, Q. Dong, Q. Wang, Y. Yuan, C. Wang, Y. Gao and J. Huang,
Energy Environ. Sci., 2014, 7, 2619.
H. Zhou, Q. Chen and Y. Yang, MRS Bulletin, 2015, 40, 667-673.
Q. Chen, H. Zhou, Z. Hong, S. Luo, H. S. Duan, H. H. Wang, Y. Liu, G. Li and Y. Yang,
J. Am. Chem. Soc., 2014, 136, 622-625.
J. Karpinska, A. Erxleben and P. McArdle, Cryst. Growth Des., 2013, 13, 1122-1130.
A. R. Mcghie, A. F. Garito and A. J. Heeger, J. Cryst. Growth, 1974, 22, 295-297.
O. Milton, Materials Science of Thin Films - Deposition and Structure Academic Press,
2002.
Y. Peng, G. Jing and T. Cui, J. Mater. Chem. A, 2015, 3, 12436-12442.
M. R. Leyden, L. K. Ono, S. R. Raga, Y. Kato, S. Wang and Y. Qi, J. Mater. Chem. A,
2014, 2, 18742-18745.
J. Mao, H. Zhang, H. He, H. Lu, F. Xie, D. Zhang, K. S. Wong and W. C. H. Choy, RSC
Adv., 2015, 5, 73760-73766.
P. Luo, Z. Liu, W. Xia, C. Yuan, J. Cheng and Y. Lu, ACS Appl. Mater. Interfaces, 2015,
7, 2708-2714.
H.-S. Ko, J.-W. Lee and N.-G. Park, J. Mater. Chem. A, 2015, 3, 8808-8815.
L. K. Ono, M. R. Leyden, S. Wang and Y. Qi, J. Mater. Chem. A, 2016.

55

Chapter 4 Characterization of perovskite solar cells by electrochemical


impedance spectroscopy

Abstract
Electrochemical impedance spectroscopy was employed to characterize perovskite solar cells.
The effect of perovskite as well as PEDOT:PSS layer thickness on the impedance response was
investigated. By using a simple equivalent circuit model, the high-frequency feature is attributed
to the recombination processes in the interfacial region of the perovskite and contact layers, and
the low-frequency feature arises from the polarization effects induced by the localized electric
field in the perovskite material. Under open-circuit conditions, it is the changing in the quality of
the perovskite films that contributes to different charge recombination dynamics.

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

4.1 Introduction
Although the efficiencies of hybrid perovskite solar cells have experienced the most rapid
increase in all kinds of photovoltaic technologies, many aspects regarding their operating
mechanisms are not yet fully understood. Moreover, it hinders the further improvement in the PV
performance.1 Traditionally, the characterization of solar cells is mostly by measuring the J-V
curves as well as the external quantum efficiency (EQE) under simulated AM1.5G illumination,
which provides information on power conversion efficiency and spectral response as a function of
wavelength, respectively. However, little intrinsic information about the solar cell, such as charge
transport mechanism, recombination process and accumulation of charges at selective contacts
are provided by these measurements. It is suggested that2 a frequency-dependent characterization
technique can be utilized to differentiate these ambiguous processes with different time constants.
Among them, the electrochemical impedance spectroscopy (EIS) is a powerful tool for a better
understanding of these time-domain processes. Developed in electrical engineering, EIS is widely
applied to different material systems, such as redox reaction at electrodes, coatings, lithium-ion
batteries, semiconductors, and dye-sensitized solar cells.3, 4 Recently, this technique has also been
investigated for characterizing perovskite solar cells, with a preliminary progress in
understanding the electro-hole recombination loss as well as hysteresis phenomena.5, 6
In an EIS measurement, the frequency-dependent impedance response of the electrical device can
be measured by employing a small sinusoidal modulation of AC bias voltage perturbation that is
superimposed on a DC bias voltage. After measurements, a proper equivalent circuit is built to
describe the electrochemical system of the device. By separating the electrical components of this
analog circuit, parameters such as resistance and capacitance can be analyzed as a function of DC
voltage, to interpret different electronic processes within the device. 7 A typical impedance
response of perovskite solar cells to the AC bias perturbation is illustrated in Fig. 4.1, where the
device was operating under light illumination and applied DC bias voltage. Usually, a simple
R(RC) or R(RC)(RC) equivalent circuit is used for modeling the experimental impedances, as is
shown in Fig. 4.2.5, 8, 9

Figure 4.1 3-D impedance spectrum of a typical perovskite solar cell, measured under 16 mW cm-2 illumination and
0.83 V at open circuit condition.

57

Chapter 4

The periodic voltage perturbation applied to the device should be small to keep the device in
equilibrium and the response linear. General definition of AC impedance of the electrical system
is given by

Z ( )

V ( )
Z cos( ) j sin( )
I ( )

(4.1)

Where both the current and the modulus of the impedance is shifted in phase by an angle on the
applied voltage. The electrical response of an ideal resistor is linear and independent of frequency,
therefore:

V ( )
R
I ( )

(4.2)

The current of a pure capacitor is shifted 90 degrees out of phase, and the impedance of a pure
capacitor is, therefore, imaginary and inversely proportional to frequency:

Z ( )

V ( )
1

I ( ) jC

(4.3)

For an equivalent circuit R(RC) as shown in Fig. 4.2(a) which consists of a parallel (RC) element
connected in series with a resistance, the AC impedance of the circuit is given by:

Z ( ) R0

1
1
jC
R1

R0

R1
R 1 C1
j
2
1 ( R1C1 )
1 ( R1C1 ) 2

(4.4)

In the complex plane, the impedance of the equivalent circuit presents a semicircle, as can be seen
in Fig. 4.2(c). When the frequency is zero, the impedance becomes Z = R0 + R1, represented by
the right-handed intersection of the arc with the Z axis. As the frequency increases, the current
will flow through the capacitance and the imaginary impedance is no longer zero. At very high
frequencies the current only flows through the capacitance of the RC element, and the impedance
is equal to R0, represented by the left-handed intersection of the arc with the Z axis. Moreover, at
the maximum point of semicircle, the characteristic frequency of which gives a time constant of
the parallel (RC) element:

1
RC

(4.5)

Similarly, the equivalent circuit R(RC)(RC) which consists of two parallel (RC) circuits in series,
as shown in Fig. 4.2(b), contains two time constants and can be represented by two semicircles on
the complex plan plots, see Fig. 4.2(d). However, when the time constants are close to each other,
instead of two well-separated semicircles, only one large semicircle is observed in the complex
plane. It is notable that the EIS measurement does not provide a direct measure of physical
processes within the device. Therefore, without having a good pre-knowledge of the material
system, the interpretation based on simply modeling is sometimes not reliable.4 To help in this
respect, analysis of spectra for different DC voltages and device configurations is commonly
implemented.
58

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

Figure 4.2 Equivalent circuits of (a) a parallel (R1C1) element connected in series with a resistance R0 and (b) two
parallel (R1,2C1,2) circuits in series with R0. Simulated impedance spectra, drawn in the complex plane: (c) One
semicircle, where the frequency increase from 0 to infinite, right-to-left,17 (d) Two semicircles, where (R1C1) represents
the high-frequency part, and (R2C2) accounts for the low-frequency part.

Since EIS is a very sensitive technique, an unstable device fabricated by a not optimized material
composition might lead to a misunderstanding of the physical processes within the device. The
material system under measuring should remain its original state after the external perturbation is
removed. Therefore, it is suggested that only the high-efficiency device with long-term stability
should be used for the EIS characterizing, to ensure a judicial interpretation of the impedance
response under working conditions. In this section, the EIS measurements were conducted and
analyzed by using the high-performance perovskite solar cells, which were fabricated by an
optimized material composition and processing method in our group. The effect of the layer
thickness of the perovskite and the hole transporting material PEDOT:PSS on the impedance
response are investigated in EIS measurement.

4.2 Perovskite layer thickness


By changing the molarity of the precursor solution, three devices with different perovskite layer
thicknesses (145 nm, 475 nm, and 700 nm) were prepared for EIS measurements. The devices
were made in the same batch using identical fabricating procedures and laboratory conditions,
based on the planar perovskite solar cell structure, which is Glass/ITO/PEDOT:PSS/CH3NH3PbI3xClx/[60]PCBM/Al. The cells were stored in an N2 filled glovebox, and their J-V characteristics
were measured before transferring for the EIS measurements, as is shown in Fig. 4.3(a). Under
100 mW cm-2 light illumination, the device with a perovskite thickness of 475 nm displays the
highest PCE of 12.6%, with the best Jsc, Voc, and FF of 15.7 mA cm-2, 0.989 V and 0.815,
respectively. After finishing the EIS measurement, the J-V performance of these devices was
measured for the second time, to investigate the stability of the devices and to establish an
accurate interpretation for the results. The J-V curves are shown in Appendix 2, and it is found
that the 145 nm and 475 nm cells remained stable with little drop in device performance while the
700 nm cell degraded seriously by displaying a substantial S-shape caused by poor extraction of
charge carriers. In this study, the impedance measurements were performed in the 1 MHz to 100
mHz frequency range using a 10 mV perturbation, with different light illumination intensities and
open-circuit voltages. Following Refs.3, 10, measurements are performed under open-circuit
voltages, for which there is no direct current and no spatial separation of charges in the perovskite
layer. Therefore, the device can operate under the recombination condition with less electron and
ionic movement induced by an external electric field, which essentially provides information of
59

Chapter 4

the devices at different charge densities. The typical impedance spectra of three devices in the
complex plane are shown in Fig. 4.3(b), where the hollow dots are the experimental data and the
solid lines are fitted curves by using the equivalent circuit in Fig. 4.2.
(a)

(b) 4000

475 nm, Voc = 0.82 V, I0 = 0.54 mW/cm2


700 nm, Voc = 0.81 V, I0 = 21.33 mW/cm2
3000

-4

-Z'' ()

Current density J (mA cm-2)

145 nm, Voc = 0.83 V, I0= 16.95 mW/cm2

145 nm
475 nm
700 nm

-8

3.9 kHz

2000

79 kHz
-12

25 kHz

1000

25 kHz
-16

100 Hz
100 Hz

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

Voltage (V)

1000

2000

3000

4000

5000

6000

Z'()

Figure 4.3 (a) J-V curves of perovskite devices stored in the glovebox, before EIS measurements. (b) Typical
impedance spectra in the complex plane, for the 145 nm, 475 nm, and 700 nm devices measured under 16.95, 0.54 and
21.33 mW cm-2 at Voc (~0.82 V), respectively. The hollow dots are experimental data, and solid lines are fitted by an
equivalent circuit shown in Fig. 4.2.

Under all the measured light intensities and corresponding open-circuit voltages, both the 145 nm
and 475 nm devices display one distinct semicircle in the high-frequency range and one tiny tail
in the low-frequency range. It is reported that10, 11 for planar perovskite devices, the highfrequency feature in the impedance spectrum relates to the charge separation and recombination
processes at the interface between perovskite and either contact layers (to be discussed later). In
comparison, the 700 nm perovskite solar cell displays two distinct semicircles in both the highand low-frequency range. Similarly, the high-frequency response could be related to the
recombination regime. However, the origin of such a distinct low-frequency feature which
attributes to the slow dynamic process is not clear.12, 13 Under the open-circuit condition, the
external electric field experienced by the perovskite bulk should be small. Therefore, the
polarization process in which dipolar units align in opposite to the electric field is negligible. 12
Besides, at a steady-state of Voc, no substantial direct current can go through the device, the
movement of charges is limited, and therefore, slow processes such as trap-state filling within the
perovskite band gap by charges should not largely contribute to the impedance response at the
low-frequency range.13 The specific low-f responses appearing in the 145 nm and 475 nm solar
cells are attributed to a contact issue, as explained below.
To study the origin of the slow kinetics, the frequency-dependent capacitance of the three cells
are compared in Fig. 4.4. All the capacitance values are directly converted from the impedance
data measured at a similar open-circuit voltage (at around 0.82 V, but under different light
intensities), to ensure the experimental condition where there are comparable charge densities
inside all the devices. It is observed that in the high-frequency region (f > 100 Hz), the
capacitance values of all three devices show a flat response to f, indicating that there is no buildup of excess charges within the perovskite devices at high frequencies. Also, as is shown in Fig.
4.4, the total capacitance of the devices at high frequencies seems to be dominated by the value of
geometric capacitance, calculated from5

60

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

Q 0 r A

V
d

(4.6)

Where 0 is the absolute dielectric permittivity, r is the relative permittivity (24, for perovskite
solar cell5), A is the cell area, and d is the thickness of the perovskite layer. However, at f < 100
Hz, the capacitance of all the cells instantly increase by several orders of magnitude compare to
their geometric capacitance values, among which the 700 nm cell displays the most substantial
one and, in correspondence with the dramatic increase of slow-responded charge concentrations
within the perovskite film. One major reason could be that10, 14 the different electron affinity of
two contact layers causes the band bending at the interface of perovskite layer, which leads to the
accumulation of electrons/ions and accordingly, a localized electric field in the interfacial region.
Therefore, the localized electric field might induce the polarization effects within perovskite
material and contributes to the slow response at low frequencies. Compared to the thinner device,
the 700 nm cell might possess the highest ionic defects due to fabrication issues or degradation
during the measurements, which gives rise to the significant low-frequency responses in both C-f
and impedance spectra.
10-2
700 nm, Voc = 0.81 V, I0 = 21.33 mW/cm2

10-3

475 nm, Voc = 0.82 V, I0 = 0.54 mW/cm2


145 nm, Voc = 0.83 V, I0= 16.95 mW/cm2

10-4

C (F/cm2)

10-5
10-6
10-7
10-8
10-9
10-10
10-1

100

101

102

103

104

105

106

Frequency (Hz)

Figure 4.4 Frequency-dependent capacitance (C-f) response. Measured under () 16.95 mW cm-2 and 0.83 V at open
circuit, for the 145 nm cell, () 0.54 mW cm-2 and 0.82 V at open circuit, for the 475 nm cell, () 21.33 mW cm-2 and
0.81 V at open circuit, for the 475 nm cell. The dotted lines represent the corresponding geometric capacitance,
ignoring the roughness factor5.

As is shown in Appendix 2, it is clear that the radii of all the high-frequency semicircles decrease
with increasing light intensities at the corresponding open-circuit voltages. By fitting these
impedance spectra to the equivalent circuit model, it is demonstrated that the shrinkage of the
high-frequency semicircles correlates to the decreasing in the resistance R1 while the series
resistance R0 remains constant. The variation trends of the R0 and R1 as a function of the opencircuit voltage are drawn in Fig. 4.5(b). As is already known, the rising in light illumination
intensity enlarges not only the charge densities and the photon-generated current densities, but
also the separation of hole and electron quasi-Fermi levels in perovskite material, and accordingly,
the final open-circuit voltage increases, as is given by:

Voc

2.303mkT
log Pin
q
61

(4.7)

Chapter 4

Where Pin is the incident light intensity, m is the ideality factor. On the other hand, under the
constant light illumination, a high forward bias voltage (Vapp) on the device also gives rise to the
charge densities within the perovskite film, but the generated forward bias diffusion current is in
the reverse direction of the photon-generated current density (Jsc). The diode equation is given
by13

q(V JRs ) Vapp JRs


J J sc J 0 exp app
1
nk
T
Rsh
B

(4.8)

where J0 is the reverse saturation current density, J is the output current density measured at the
cell electrodes, and n is also the ideality factor. When the series resistance Rs is minor while Rsh is
large, both parameters can be omitted and the diode equation becomes:

qV
J J sc J 0 exp app 1 J sc J rec
nk BT

(4.9)

Then the forward bias diffusion current is regarded as the recombination current Jrec and therefore,
at a larger forward bias, the recombination current increases while less photon-generated charges
can be extracted at the electrodes. Furthermore, the recombination process within the device can
be better represented by a simple electrical component, such as the voltage-dependent
recombination resistance, Rrec:
1

Rrec

qVapp
1 J rec

R0 exp
qA V
nk BT

(4.10)

Where R0 is a pre-exponential factor. A higher Vapp relates to the higher charge concentrations and
therefore, the recombination process increases which corresponds to the decrease in Rrec, resulting
in less charge collection. At the open-circuit conditions (Vapp = Voc) where the recombination is
predominant, the forward bias diffusion current equals with the Jsc (Jrec = Jsc), therefore J = 0, and
we have

qV
J 0 exp oc 1 J sc
nk BT

(4.11)

Assuming that Jsc is proportional to the light intensity, when Vapp = Voc, the comparison of
Equation (4.7) with Equation (4.10), gives that the two ideality factors should be the same, m = n.
The ideality factor can be determined by the slope of the light intensity dependence of the opencircuit voltage plots, according to

Voc
2.303mk BT

log I 0
q

(4.12)

Where I0 is the incident light intensity. As is shown in Fig. 4.5(a), the deviation from ideality
(m > 1) is observed for both the 145 nm and 700 nm perovskite cells, of which the m are 1.6 and
62

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

2.0, respectively. When m is close to 2, the recombination process arises at the interfacial region
between the perovskite film and contact layers, due to a trap-mediated process (via ionic
defects).11 As a result, under the same light illumination condition, the open-circuit voltages for
both cells are much smaller than the Voc of the 475 nm perovskite cell, which shows an ideality
factor m (= 1.1) close to 1. This denotes a band to band recombination mechanism of free carriers,
with no substantial trapping contribution.
(b) 10

(a)

Rs, 145 nm

0.9

Rs, 475 nm

107

Rs, 700 nm
R1, 145 nm, n = 1.7

10

0.8

R1, 475 nm, n = 1.4


R1, 700 nm, n = 3.1

0.7

R ()

Voc (V)

105
104
103

0.6

102
0.5

0.4
0.1

145 nm, m = 1.6


475 nm, m = 1.1
700 nm, m = 2.0
1

10

Light intensity (mW cm-2)

101
100

100

0.0

0.2

0.4

0.6

0.8

1.0

Voc (V)

Figure 4.5 (a) Light intensity dependence of Voc and (b) fitted resistance R0, R1 as a function of Voc for () 145 nm, ()
475 nm, and () 700 nm perovskite solar cells.

Moreover, by determining the slope of the variation of the recombination resistance with opencircuit voltage, the ideality factor n can also be obtained, according to

log Rrec
q

Voc
2.303nk BT

(4.13)

After fitting the equivalent circuit to the impedance spectra, the high-frequency resistance R1 is
plotted as a function of Voc, shown in Fig. 4.5(b). All the three semi-logarithmic R1 plots show the
exponential dependence on the open-circuit voltage. (The first two R1 data points of the 175 nm
cell at low light illumination are omitted, since the effect of shunt resistance is large which leads
to the possibility of a leak current, and the variation trend is no longer in exponential form.) The
ideality factors n of the 145 nm and 475 nm perovskite cells are 1.7 and 1.4, respectively. The
ideality factors obtained from these slopes are summarized in Table 4.1. Despite some differences,
it is reasonable to propose that the n, calculated from R1, is very similar to the m from Fig. 4.5(a).
Therefore, it is concluded that the electrical component R1 can represent the recombination
resistance within the perovskite devices. For the 700 nm perovskite cell, however, the obtained
ideality factor n of 3.1, is much larger than the m of 2.0. This could be attributed to the
degradation of the perovskite cell during the EIS measurements, as is evidenced by the unique Sshape of the J-V curve, shown in Appendix 2. The increased (ionic) defects, therefore, act as traps
for carriers, which give rise to the recombination processes in the devices. In conclusion, the
decent PV performance (higher Voc and FF) of the 475 nm perovskite solar cell is consistent with
its small ideality factors (m = 1.1, n = 1.4), which denotes a different recombination mechanism,
and leads to a larger recombination resistance. Hence, under the same light illumination level, the
photon-generated charge density within the 475 nm cell is the largest and results in the highest
open-circuit voltage for the three cells.

63

Chapter 4

Table 4.1 Ideality Factors m and n, derived from the slope in Fig. 4.5(a) and (b), respectively.

Ideality factor
m
n

145 nm
1.6
1.7

475 nm
1.1
1.4

700 nm
2.0
3.1

It is remarkable that, including the 145 nm and 475 nm devices, the ideality factor n calculated
from Fig. 4.5(b) is always higher than the m, which can be caused by many reasons. As is
discussed above, the most likely reason is the degradation of the perovskite device during the EIS
measurements, by either the moisture that oxidizes the Al electrode,15 or the decomposition of
active layer under a long-term light illumination, which arises the trap-assisted recombination
processes. Also, the precise determination of the steady-state at Voc for each light intensity is
difficult. Therefore, the movement of charges via micro-current might also contribute to
impedance response at high-frequency region and results in the deviated resistance values. Lastly,
the R(RC) equivalent circuit used for EIS fitting might not be the optimum choice to interpret the
spectral features. There have been several different electrical models developed for characterizing
planar perovskite solar cells in EIS,5, 10, 16 but the validity of these models remains further
investigations.

4.3 PEDOT:PSS layer thickness


By changing the spin-coating speeds, perovskite devices with different hole-transporting layer
thicknesses (PEDOT:PSS, 25 nm, 40 nm and 60 nm) were fabricated, while the thickness of
perovskite layer kept at 470 nm. Under identical laboratory conditions, all the cells were
fabricated in one batch based on the planar perovskite device structure: Glass/ITO/PEDOT:
PSS/CH3NH3PbI3-xClx/[60]PCBM/Al. Fig. 4.6(a) shows the J-V curves of three solar cells stored
in an N2 filled glovebox, before transferring for the EIS measurements. Under 100 mW cm-2 light
illumination, the device with 25 nm PEDOT:PSS layer displays the best J-V performance, with a
PCE of 15.0% in the fast, downward sweep, yielding a Jsc of 19.1 mA cm-2, Voc of 0.993 V and FF
of 0.789. After the long-term EIS measurement, the J-V characteristics of 25 nm and 40 nm
devices remain unchanged, while the 60 nm device displays peculiar S-shape in the J-V curves, as
is shown in Appendix 3. The EIS measurements were performed at 1 MHz to 100 mHz frequency
range using a 10 mV AC perturbation while the devices were operated at different light
illumination intensities and open-circuit voltages. Fig. 4.6(b) displays the typical impedance
spectra of three devices in the complex impedance plane, in which the hollow dots are the
experimental data, and the solid lines are the equivalent circuit fits.

64

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

300

(a) 4

25 nm
40 nm
60 nm

40 nm, Voc = 0.93 V, I0 = 61.52 mW/cm2

250

60 nm, Voc = 0.93 V, I0 = 61.52 mW/cm2

50 kHz

200

-4

-Z'' ()

Current density J (mA cm-2)

25 nm, Voc = 0.93 V, I0= 16.95 mW/cm2

-8

150

-12

100

-16

50

-20

125 kHz

100 Hz
-0.2

0.0

0.2

0.4

0.6

0.8

158 kHz

1.0

100

200

Voltage (V)

300

400

500

Z' ()

Figure 4.6 (a) J-V curves of perovskite devices, under 100 mW cm-2, before EIS measurements. (b) Typical impedance
spectra in the complex plane, for the devices fabricated from 25 nm, 40 nm, and 60 nm thick PEDOT:PSS layer,
measured under 16.95, 61.52 and 61.52 mW cm-2 at Voc = 0.93 V, respectively. The hollow dots are the experimental
data, and solid lines are fitted by an equivalent circuit in Fig. 4.2.
(b) 10

(a) 1.00

Rs, 25 nm
Rs, 40 nm

0.95

Rs, 60 nm
R1, 25 nm, n = 1.4
R1, 40 nm, n = 1.1

103

0.90

R ()

Voc (V)

R1, 60 nm, n = 1.2


0.85

102

0.80

25 nm, m = 1.0
40 nm, m = 1.0
60 nm, m = 1.0

0.75

101
0.82

0.70
1

10

100

0.84

0.86

0.88

0.90

0.92

0.94

0.96

0.98

Voc (V)

-2

Light intensity (mW cm )


(c) 1E-5

25 nm
40 nm
60 nm

C (Fcm-2)

1E-6

1E-7

1E-8

1E-9
0.82

0.84

0.86

0.88

0.90

0.92

0.94

0.96

0.98

Voc (V)

Figure 4.7 (a) Light intensity dependence of Voc, (b) fitted resistance R0, R1 as a function of Voc, and (c) fitted
capacitance C1 as a function of Voc for perovskite solar cells fabricated from () 25 nm, () 40 nm, and () 60 nm
PEDOT:PSS layer. The dotted line represents the corresponding geometric capacitance, ignoring the roughness factor.

Under different light illumination levels and Voc, the impedance responses of all three devices
display one distinct semicircle in the high-frequency range and one peculiar tail in the lowfrequency range (f < 100 Hz). The low-f response could be ascribed to the deviation from the
steady state of Voc, and therefore, a localized electric field remains in the perovskite layer and
leads to the alignment of dipole units in response to the perturbation.10 As shown in Appendix 3,
it is observed that the magnitude of the semicircle at high-frequency (f > 100 Hz) decreases with
65

Chapter 4

the increasing light intensities and Voc, corresponding to the exponential decreasing in R1 as a
function of Voc, as can be seen in the semi-logarithmic plot of R1 in Fig. 4.7(b). Similar to the
discussions in Section 4.2, the recombination dynamics of perovskite solar cells is examined
either by plotting the light intensity dependence of Voc as is shown in Fig. 4.7(a), or by showing
the fitted high-frequency resistance R1 as a function of Voc in Fig. 4.7(b), provided that the
ideality factors calculated in both cases be comparable (m = n). By linearly fitting these plots, the
ideality factors obtained from the slopes are summarized in Table 4.2. Although displaying a
difference in ideality factors of the 25 nm cell (where m = 1.0 and n = 1.4), it is still reasonable to
claim that the electrical component R1 represents the recombination resistance of the perovskite
devices. Therefore, the fact that the ideality factors close to 1 are held almost entirely true for all
three perovskite devices (except for the n of the 25 nm cell), which demonstrates that the change
in PEDOT:PSS layer thicknesses do not contribute to different recombination regimes. Compared
to the m and n in Section 4.2, it is the quality of the perovskite layer that regulates the
recombination dynamics (slope) of the solar cell while the PEDOT:PSS layer thickness only
affect the absolute value of the Rrec. The 25 nm cell shows the highest Rrec among all three cells,
which relates to less charge recombination and, therefore, the higher Voc and FF in the J-V curves.
Table 4.2 Ideality factors m and n, derived from the slope in Fig. 4.7(a) and Fig. 4.7(b), respectively.

Ideality factor
m
n

25 nm
1.0
1.4

40 nm
1.0
1.1

60 nm
1.0
1.2

After fitting the high-frequency impedance spectra, for all three cells the capacitance C1, which is
in parallel with the R1 of the equivalent circuit, shows a flat response to the Voc, as shown in Fig.
4.7(c). The phenomenon that the magnitude of the high-frequency capacitance does not increase
with light intensities is similar to the Helmholtz response of an electrical double layer.10 C1 only
reflects the charges stored in the contacts between perovskite and transporting layers where they
show a linear variation trend with voltage. In the meantime, the amount of photon-generated
charges stored inside the perovskite layer is relatively small and do not help to increase the C1.
Therefore, the perovskite layer can be treated as a low photon-doped intrinsic layer. According to
Equation (4.6), Cgeo is calculated and plotted in Fig 4.7(c), which ignores the roughness factor for
perovskite film (ca. 2~5)5 and hence it is several times smaller than the actual geometric
capacitance, implying that the fitted C1 is consistent with the real Cgeo values. In conclusion, by
fitting the high-frequency capacitance of the solar cells, we cannot obtain information about the
charge concentrations inside the perovskite layer, since the charge densities in the contacts are
much larger than the photon-generated charges inside the perovskite layer.

4.4 Conclusions
To obtain a deeper understanding of the time-domain charge transporting and recombination
mechanisms of the perovskite solar cells, a frequency-dependent characterization technique
named electrochemical impedance spectroscopy (EIS) measurements were conducted, by using
the lab-produced high-performing perovskite devices. All the experiments were taken under 1
MHz to 100 mHz frequency range using a 10 mV perturbation, with different light illumination
intensities and corresponding open-circuit voltages. Firstly, the effect of the perovskite layer
thickness on the impedance response was investigated. In the complex plane of all three cells, the
700 nm cell displays a distinct low-f response, which could be attributed to the accumulation of
charges and a localized electric field at the perovskite interfacial region, inducing the polarization
66

Characterization of perovskite solar cells by electrochemical impedance spectroscopy

of dipole units and response to the low-frequency perturbation. By comparing the ideality factors
obtained from both the light intensity dependent open-circuit plot and the plot of the highfrequency resistance R1 as a function of Voc, it is concluded that R1 in the equivalent circuit
represents the recombination resistance of the perovskite devices. The 475 nm cell displays the
smallest ideality factor and the largest recombination resistance, in correspondence with a higher
charge density under the same light illumination level, and results in the highest open-circuit
voltage of all three cells. Secondly, it is the perovskite layer that regulates the recombination
dynamics, while the change in PEDOT:PSS layer thickness do not contribute to different
recombination mechanisms. Moreover, the device with the thinnest PEDOT:PSS layer gives the
best PV performance, due to a larger recombination resistance.

4.5 Experiments
An electrochemical station (Solartron, SI 1260) and a data acquisition software (SMaRT v3.3.1)
were used. The incident light source was generated by using a 50 W tungsten-halogen lamp
(Muller, LXH 100), calibrated by a Si reference cell. The light intensities were controlled by
tuning the two neutral-density filters. To avoid degradation in air, perovskite device was sealed in
a spectral response box, before measuring the impedance spectroscopy. The devices were
measured under open-circuit conditions by illuminating the devices at different light intensities.
The active area of the cell was 0.09 cm2, without using a mask. By applying a 10 mV AC
perturbation to the device, the impedance spectra were recorded in two steps. Firstly, the
frequency changed from 1 MHz to 100 Hz in 100 steps, with an integration period of 10 seconds
per point; Secondly, to ensure a precise low-f measurement, 100 points were recorded from 100
Hz to 0.1 Hz, and each point is averaged over 100 AC cycles. Software (Zview 2) was used for
analyzing the obtained impedance spectra.

67

Chapter 4

4.6 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

E. Guilln, F. J. Ramos, J. A. Anta and S. Ahmad, J. Phys. Chem. C, 2014, 118, 2291322922.
A. Zohar, N. Kedem, I. Levine, D. Zohar, A. Vilan, D. Ehre, G. Hodes and D. Cahen, J.
Phys. Chem. Lett., 2016, 7, 191-197.
F. Fabregat-Santiago, G. Garcia-Belmonte, I. Mora-Sero and J. Bisquert, Phys. Chem.
Chem. Phys., 2011, 13, 9083-9118.
A. Lasia, Electrochemical Impedance Spectroscopy and its Applications, Springer,
Springer New York, 2014.
A. Pockett, G. E. Eperon, T. Peltola, H. J. Snaith, A. Walker, L. M. Peter and P. J.
Cameron, J. Phys. Chem. C, 2015, 119, 3456-3465.
H. S. Kim, I. H. Jang, N. Ahn, M. Choi, A. Guerrero, J. Bisquert and N. G. Park, J. Phys.
Chem. Lett., 2015, 6, 4633-4639.
A. Todinova, J. Idigoras, M. Salado, S. Kazim and J. A. Anta, J. Phys. Chem. Lett., 2015,
6, 3923-3930.
T. Du, N. Wang, H. Chen, H. Lin and H. He, ACS Appl. Mater. Interfaces, 2015, 7, 33823388.
D. Liu, M. K. Gangishetty and T. L. Kelly, J. Mater. Chem. A, 2014, 2, 19873-19881.
A. R. Pascoe, N. W. Duffy, A. D. Scully, F. Huang and Y.-B. Cheng, J. Phys. Chem. C,
2015, 119, 4444-4453.
J. Liu, S. Pathak, T. Stergiopoulos, T. Leijtens, K. Wojciechowski, S. Schumann, N.
Kausch-Busies and H. J. Snaith, J. Phys. Chem. Lett., 2015, 6, 1666-1673.
R. S. Sanchez, V. Gonzalez-Pedro, J. W. Lee, N. G. Park, Y. S. Kang, I. Mora-Sero and J.
Bisquert, J. Phys. Chem. Lett., 2014, 5, 2357-2363.
K. Miyano, N. Tripathi, M. Yanagida and Y. Shirai, Acc. Chem. Res., 2016, 49, 303-310.
I. Zarazua, J. Bisquert and G. Garcia-Belmonte, J. Phys. Chem. Lett., 2016, 7, 525-528.
Y. Han, S. Meyer, Y. Dkhissi, K. Weber, J. M. Pringle, U. Bach, L. Spiccia and Y.-B.
Cheng, J. Mater. Chem. A, 2015, 3, 8139-8147.
V. Gonzalez-Pedro, E. J. Juarez-Perez, W. S. Arsyad, E. M. Barea, F. Fabregat-Santiago,
I. Mora-Sero and J. Bisquert, Nano Lett., 2014, 14, 888-893.

68

Summary
In this report, a simplified and easy-manipulated sequential deposition approach was investigated
for the fabrication of planar perovskite solar cells in the regular device configuration (p-i-n).
Several kinds of lead halide precursor films were employed to react with methylammonium
iodide both in solution and vapor phase, to seek an efficient strategy to produce the perovskite
absorber. It was found that the compositions and morphologies of the lead halide precursor films
play a conspicuous role in controlling the growth of perovskite films. Also, the manner that the
CH3NH3I deposited onto the inorganic substrates has a significant impact on the final device
performances.
First of all, a c-PbI2 (compact, 140 nm) film was found to have serious blocking issues from the
reaction with different concentrations of CH3NH3I solutions, by rapidly producing a compact and
volume-expanded perovskite layer on the surface, which isolated the bottom PbI2 from the
conversion into perovskite phase, and led to an insufficient active layer thickness. Simply
increase the reaction time between the two components resulted in a non-stoichiometric condition
or the dissolution of perovskite, and no improvement in efficiency was achieved. After
optimization, an efficiency of 7.52% was obtained, with Jsc of only 10.4 mA cm-2, in correlation
to a perovskite thickness of 232 nm. By contrast, the porous PbI2 (140 nm) facilitated the
penetration of organic solutions, which displayed a much faster conversion rate with CH 3NH3I,
by showing a high perovskite absorption after a 20 s dipping procedure, compared to the weak
absorption obtained from the c-PbI2 film. After optimizing the annealing process, the best device
exhibited an efficiency of 8.95%, with an increased Jsc of 13.78 mA cm-2, and a perovskite layer
thickness of 262 nm. On the other hand, it was found that the incorporation of PbCl 2 in the
precursor solution promoted the conversion into perovskite by releasing volatile organic
byproducts. However, a large concentration of PbCl2 would lead to a drastic evaporation of
gaseous byproducts, which deteriorated the device performance. The mixed halide precursor
films (165 nm) displayed a slightly slower conversion than the p-PbI2 films. After a 60 s hot
dipping process, the optimized device gave an efficiency of 8.04%, with a Jsc of 13.40 mA cm-2
and a perovskite film thickness of 259 nm.
It was found that during the vapor assisted deposition process, the substrate temperature was the
most critical parameter that determined the growth of the perovskite film, by controlling a
balance between the physical vapor deposition rate and chemical conversion rate. In a thermal
gradient sublimer, the achievable substrate temperature was way lower than the evaporation
temperature, then the chemical reaction between the lead halide precursor film and CH 3NH3I was
therefore much slower than the physical vapor deposition process. After a long-term evaporation,
the over-rapid deposition of CH3NH3I resulted in a very rough surface, while the diffusion of the
organic phase was severely hindered and led to a limited conversion to perovskite. The device
obtained from the c-PbI2 precursor (160 nm) displayed an efficiency of 7.65% with the perovskite
thickness of 269 nm. By using the p-PbI2 substrate (145 nm), a more homogeneous vapor
deposition process was achieved. However, the conversion was not improved, with a resultant
device efficiency of only 7.04% (perovskite thickness of 210 nm). To achieve an efficient vapor
deposition process, a home-designed two-zone furnace was built, with better control over the
temperature parameters, as well as the atmosphere involved in the formation of perovskite film.
In the EIS measurement, the low-frequency impedance response was mainly attributed to the
ionic polarization effect induced by the localized electric field built at the interfacial region.
69

Meanwhile, the high-frequency impedance response was found related to the recombination
process at the interface between perovskite and either contact. Due to fabrication issues or the
degradation of devices during the EIS measurements, a thick perovskite film (700 nm) was found
to possess a high density of ionic defects, which was evidenced by the significant low-frequency
response in both C-f and impedance spectrum. Furthermore, the increased ionic defects gave rise
to the trap-assisted recombination processes and corresponded to a smaller recombination
resistance derived from the high-frequency response. Under the same light illumination condition,
it was demonstrated that a poorly fabricated perovskite device would display a low Voc and FF,
consistent with its low charge concentrations in the perovskite layer.

70

Appendix 1
The home-built VASP setup

220

220

200

200

180

180

160

160

140
T1: 140 C
T2: 180 C
T5: 180 C

120
100
80

T1: 140 C
T2: 160 C
T5: 180 C

T1: 140 C
T2: 140 C
T1:
T5: 180 C

120 C
T2: 140 C
T5: 180 C

60
40

0
0

1000

2000

3000

4000

5000

6000

7000

140
120
100
80
T1: 120 C
T5: 180 C

60

Temp - T1 Sample holder


Temp - T2 Crucible
Temp - T3 Inner wall
Temp - T4 Inner wall
Temp - T5 Heating ring

T1: 140 C
T5: 180 C

20

Temperature (C)

Temperature (C)

Figure 1 The photograph of the home-built two-zone furnace for vapor phase deposition.

40

T1: 140 C
T1: 140 C T2: 180 C
T2: 160 C T5: 180 C
T1: 120 C
T5: 180 C
T2: 140 C
T5: 180 C
Temp - T1 Sample holder
Temp - T2 Crucible
Temp - T3 Inner wall
Temp - T4 Inner wall
Temp - T5 Heating ring

T5: 180 C

20
0
8000

1000

2000

3000

Time (sec)

4000

5000

Time (sec)

Figure 2 Temperature-time graphs of the device.

71

6000

7000

8000

9000

Appendix 2
J-V and impedance spectra of 145 nm, 475 nm and 700 nm (perovskite layer thickness)
perovskite solar cells.
4

(b)

(a)
2

Before, downward
Before, upward

Current density J (mA cm-2)

Current density J (mA cm-2)

0
-2
-4
-6
-8
-10

After, downward
After, upward

0
-2
-4
-6
-8
-10
-12

-12
-0.2

0.0

0.2

0.4

0.6

0.8

1.0

-0.2

0.0

0.2

Voltage (V)

0.6

0.8

1.0

(d) 4

(c) 4
Before, backward
Before, forward

Current density J (mA cm-2)

Current density J (mA cm-2)

0.4

Voltage (V)

-4

-8

-12

After, backward
After, forward

-4

-8

-12

-16
-0.2

0.0

0.2

0.4

0.6

0.8

-16
-0.2

1.0

0.0

0.2

Voltage (V)
(e)

(f)

Current density J (mA cm-2)

Current density J (mA cm-2)

Before, downward

-4

-8

-12

-0.2

0.6

0.8

1.0

After, downward
After, upward

0
0

0.4

Voltage (V)

-2
-4
-6
-8
-10
-12

0.0

0.2

0.4

0.6

-14
-0.2

0.8

Voltage (V)

0.0

0.2

0.4

0.6

0.8

1.0

Voltage (V)

Figure 3 Under 100 mW cm-2 illumination, J-V curves of the (a) 145 nm, (c) 475 nm, and (e) 700 nm thick perovskite
film measured before and after (b) 145 nm, (d) 475 nm, and (f) 700 nm the EIS measurements.

72

Table 1 Under 100 mW cm-2 illumination, PV parameters of 145 nm, 475 nm and 700 nm cells measured in downward
direction, before and after EIS measurements.

Downward
Direction
Before
After
Before
After
Before
After

Thickness
145 nm
475 nm
700 nm
(a)

Jsc (mA cm-2)

Voc (V)

FF

PCE(%)

10.7
11.5
15.7
14.3
12.9
11.9

0.891
0.919
0.989
0.981
0.810
0.802

0.537
0.573
0.815
0.775
0.756
0.286

5.14
6.04
12.6
10.9
7.89
2.72

(b) 2500

Voc = 0 V, I0 = 0 mW/cm2

80000

Voc = 0.75 V, I0 = 6.75 mW/cm2

Voc = 0.48 V, I0 = 0.13 mW/cm2


Voc = 0.65 V, I0 = 0.54 mW/cm

70000

Voc = 0.83 V, I0 = 16.95 mW/cm2

Voc = 0.84 V, I0 = 38.81 mW/cm2

Voc = 0.71 V, I0 = 1.35 mW/cm2

Voc = 0.86 V, I0 = 61.52 mW/cm2

2000

Voc = 0.88 V, I0 = 97.51 mW/cm2

1500

50000

-Z'' ()

-Z'' ()

60000

40000

1000

30000
20000

500

10000
0

0
0

50000

100000

150000

200000

1000

2000

3000

Z' ()
(c)

Voc = 0.78 V, I0 = 0.13 mW/cm2

7000

(d)

Voc = 0.87 V, I0 = 1.35 mW/cm

6000

Voc = 0.59 V, I0 = 0.54 mW/cm2

Voc = 0.61 V, I0 = 1.35 mW/cm2

12000

Voc = 0.89 V, I0 = 6.75 mW/cm2

Voc = 0.67 V, I0 = 6.75 mW/cm2

Voc = 0.92 V, I0 = 16.95 mW/cm2

Voc = 0.78 V, I0 = 16.95 mW/cm2

10000

5000

-Z''()

4000

-Z''()

5000

Voc = 0.54 V, I0 = 0.13 mW/cm2

14000

Voc = 0.82 V, I0 = 0.54 mW/cm2

6000

4000

Z'()

3000

Voc = 0.81 V, I0 = 21.33 mW/cm2

8000
6000
4000

2000

2000

1000

0
0

4000

8000

12000

16000

10000

20000

30000

40000

50000

Z'()

Z' ()

Figure 4 Impedance spectra of (a)(b) 145 nm, (c) 475 nm, and (d) 700 nm devices measured under different light
intensity at Voc. The hollow dots are the experimental data, and solid lines are fitted by the equivalent circuit shown in
Fig. 4.2.

73

Appendix 3
J-V and impedance spectra of the 25 nm, 40 nm and 60 nm (PEDOT:PSS layer thickness)
perovskite solar cells.
(a) 4

(b) 4

Before, downward (dark)


Before, upward (dark)
Before, downward
Before, upward

-4

Current density J (mA cm-2)

Current density J (mA cm-2)

-8

-12

-16

After, downward (dark)


After, upward (dark)
After, downward
After, upward

-4

-8

-12

-16

-20
-0.2

0.0

0.2

0.4

0.6

0.8

-20
-0.2

1.0

0.0

0.2

Voltage (V)

(c)

Before, upward (dark)


Before, downward (dark)
Before, upward
Before, downward

-4

Current density J (mA cm-2)

Current density J (mA cm-2)

0.8

1.0

-8

-12

-16

0.0

0.2

0.4

0.6

0.8

0.8

1.0

After, downward (dark)


After, upward (dark)
After, downward
After, upward (dark)

-4

-8

-12

-16

-20
-0.2

1.0

0.0

0.2

0.4

0.6

Voltage (V)

Voltage (V)
(e) 4

(f) 4

Before, downward (dark)


Before, upward (dark)
Before, downward
Before, upward

-4

Current density J (mA cm-2)

Current density J (mA cm-2)

0.6

(d) 4

-20
-0.2

0.4

Voltage (V)

-8

-12

After, downward (dark)


After, upward (dark)
After, downward
After, upward

-4

-8

-12

-16
-0.2

0.0

0.2

0.4

0.6

0.8

-16
-0.2

1.0

0.0

0.2

0.4

0.6

0.8

1.0

Voltage (V)

Voltage (V)

Figure 5 Under 100 mW cm-2 illumination, J-V curves of the (a) 25 nm, (c) 40 nm, and (e) 60 nm (PEDOT:PSS
thickness) perovskite cells measured before and after (b) 25 nm, (d) 40 nm, and (f) 60 nm the EIS measurement.

74

Table 2 Under 100 mW cm-2 illumination, PV parameters of 145 nm, 475 nm and 700 nm cells measured in downward
direction, before and after EIS measurements.

Downward
Direction
Before
After
Before
After
Before
After

Thickness
25 nm
40 nm
60 nm
(a) 800

Jsc (mA cm-2)

Voc (V)

FF

PCE(%)

19.1
18.6
19.0
19.0
16.0
13.3

0.993
0.977
0.964
0.951
0.966
0.928

0.789
0.782
0.817
0.821
0.853
0.493

15.0
14.2
14.9
14.9
13.2
6.06

(b) 1200

Voc = 0.87 V, I0 = 6.75 mW/cm2

Voc = 0.83 V, I0 = 1.35 mW/cm2


Voc = 0.87 V, I0 = 6.75 mW/cm2

Voc = 0.91 V, I0 = 10.69 mW/cm2

Voc = 0.90 V, I0 = 16.95 mW/cm2

Voc = 0.93 V, I0 = 16.95 mW/cm2

700

Voc = 0.97 V, I0 = 97.51 mW/cm

600

Voc = 0.91 V, I0 = 26.86 mW/cm2

1000

Voc = 0.95 V, I0 = 48.87 mW/cm2

Voc = 0.87 V, I0 = 61.52 mW/cm2

800

400

-Z''()

-Z''()

500

300

600

400

200

200
100

0
0

200

400

600

800

1000

1200

1400

1600

1000

2000

Z'()

Z'()
(c) 500

Voc = 0.87 V, I0 = 6.75 mW/cm

Voc = 0.89 V, I0 = 10.69 mW/cm2


Voc = 0.90 V, I0 = 48.87 mW/cm2
Voc = 0.93 V, I0 = 61.52 mW/cm2

400

Voc = 0.95 V, I0 = 97.51 mW/cm2

-Z''()

300

200

100

0
0

250

500

750

1000

Z'()

Figure 6 Impedance spectra of devices with (a) 25 nm, (b) 40 nm, and (c) 60 nm PEDOT:PSS thicknesses measured
under different light intensity at Voc. The hollow dots are the experimental data, and solid lines are fitted by the
equivalent circuit shown in Fig. 4.2.

75

Acknowledgements
First and foremost, I would like to thank my supervisor Bardo Bruijnaers. It has been a great
honor to work with someone that performs on such a high level in research. Ever since the first
batch of perovskite solar cells that I made under your guidance (despite the performance was
pretty bad), I have learned an immense amount of things from you. The door to your office is
always open whenever I have a question (some of them were stupid), and you have always been
so patient to explain things to me. You taught me how to question thoughts and express ideas.
This report would not hot have been possibly finished without your help. I wish you a lot of
success in your future Ph.D. research work and a very happy life.
Furthermore, I want to thank Professor Rene Janssen for giving me the opportunity to do my
graduation project on such an appealing research topic. I feel very lucky to work in such a wellregarded group and be able to explore on my own. Many thanks for your inspiring conversations
during the perovskite group meetings, and for giving me advice when the project did not go as
originally planned.
I would also like to thank Dr. Martijn Wienk, Dr. Koen Hendriks and Hans van Franeker for their
invaluable suggestions about my project and for helping me all kinds of issues in the device lab.
Dr. Stefen Meskers, thank you for helping me arranging the whole setup for the impedance
measurements which makes my life much easier. Your teaching style has inspired me a lot, and I
enjoy all the discussions with you.
I want to express my special thanks to the visiting Professor Juan Antonio Anta, thank you for
your support during the last two months of my graduation project, and for teaching me so many
things about characterizing solar cells by impedance spectroscopy. You are always so kind to
answer all my questions about how to interpret the data. I am thankful for your valuable advice
for the experimental design and for commenting on my views in Chapter 4. It is a pleasure to
meet and talk with you.
Lastly, I would like to express my heartfelt gratitude to my parents and my loving girlfriend.
Thank you for your understanding and for giving me so much freedom to make my choices.
Living and studying alone in a foreign country is not always easy to me. None of this would have
been possible without your love, support and continuous encouragement. I wish I can go back and
see you soon.

76

You might also like