80
Materials interface engineering in perovskite photovoltaics by Jixian XU A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy The Edward S. Rogers Sr. Department of Electrical & Computer Engineering University of Toronto © Copyright by Jixian XU 2017

Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

  • Upload
    others

  • View
    18

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

Materials interface engineering in perovskite photovoltaics

by

Jixian XU

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

The Edward S. Rogers Sr. Department of Electrical & Computer Engineering

University of Toronto

© Copyright by Jixian XU 2017

Page 2: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

ii

Materials interface engineering in perovskite photovoltaics

Jixian XU

Doctor of Philosophy

The Edward S. Rogers Sr.

Department of Electrical & Computer Engineering »

University of Toronto

2017

Abstract

Solar photovoltaics (PV) offer a sustainable solution to the daunting challenge of meeting

the global energy demand. Perovskite solar cells, whose high efficiencies are attainable via

low-cost and high-throughput solution processing, are an emerging technology that has

captivated the PV research community. Further advances in efficiency are limited by the

abundant interfaces that make up these polycrystalline devices. Important issues in

perovskite device operation, such as instability and hysteresis, arise from perovskites’ ionic

nature, and need to be addressed for this technology to fulfill its potential.

In this thesis, I explore interfaces within perovskite devices: grain boundaries, and

electron- and hole-extraction junctions. With the aid of density functional theory (DFT)

simulations and nano-probe characterization, I provide insight into the origins of defect

formation and hysteresis. By leveraging these findings, I demonstrate control of film

growth conditions and interface materials chemistry to create new device architectures

with improved performance. The DFT-based analysis of defect formation energies

identifies the key defects (Pb atom substituted by I, known as antisites) and indicates that

films grown under iodine-rich conditions are prone to forming deep electronic traps. This

Page 3: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

iii

finding motivated my exploration of a new precursor (anhydrate lead acetate) for device-

quality films.

I then report the first perovskite-PCBM hybrid solid. Here, I find that PCBM, when it

infiltrates throughout the grain boundaries and electron-extraction interfaces, suppresses

hysteresis in devices. Materials characterization and DFT simulations reveal the PCBM-

perovskite interaction: the PCBM passivates the key defects during the perovskite self-

assembly. Using conductive AFM, I reveal the memristive properties of perovskite films

and identify the major origin of hysteresis as ion, especially halide, migration.

I close by developing the first crosslinked hole-extraction top contact with the goal of

obviating degradation of the underlying perovskite. A remote-doping strategy introduces

the needed hole conductivity. The new top contact produces an insoluble and heat-resistant

protecting interlayer that is band-aligned with the perovskite. The resultant family of

devices is hysteresis-free, with fill factors exceeding 80% and resilience to thermal stresses

that exceed 100˚C, conditions under which conventionally-contacted devices fail. This top

contact methodology also paves the way for building multi-junction devices on top of the

perovskite cell. I close this work by offering a roadmap for future improvements in

perovskite photovoltaics.

Page 4: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

iv

Acknowledgments

Firstly, I give my sincere thanks to my supervisor, Prof. Ted Sargent, for providing me with the

opportunity to work with fantastic colleagues and for guiding me throughout my doctoral work.

His passion and broad view of research encouraged me to solve problems across multi-

disciplinary fields. His ambitions to address world-class challenges inspired me to realize my

potential.

I also thank all the professors that have spent time on my courses and committees, and

collaborated with me in my research. I would like to acknowledge the support from the Edward

Rogers Sr. Graduate Scholarship and the Hatch Graduate Scholarship.

I feel so lucky to have worked with remarkable people through these years. A big thanks to

Andrei Buin and Oleksandr Voznyy for collaboration on DFT and optoelectronic simulations and

related publications. I give special thanks to Alex Ip, Brandon Sutherland and Grant Walters for

close collaboration in projects and patience with manuscript revision. I would like to thank Dr.

Zhijun Ning for ramping me up when I first started in the group. I also would like to thank Dr.

Sjoerd Hoogland, Larissa Levina, Elenita Palmiano, Damir Kopilovic, and Remi Wolowiec for

their continuous assistance with key aspects of my research through these years. As well, I would

like to thank Wei Li, Riccardo Comin, Xinzheng Lan, and Mingjian Yuan for their insights on

chemistry and physics. Thanks to the unsung heroes that keep the group running, Jeannie and

Stacy. Thanks and best of luck to those I had a chance to discuss and work alongside, Valerio

Adinolfi, Chris Wong, André Labelle, Susanna Thon, Illan Kramer, David Zhitomirsky,

Pongsakorn Kanjanaboos, Jeffrey McDowell, Lisa, Xiwen Gong, Mengxia Liu, Lina Quan, Min

Liu, Bo Zhang, Fengjia Fan, Zhenyu Yang, and Haopeng Dong. Although I won’t be able to

name all of them here, I am grateful for contributions from everyone.

I am grateful to my family, Mom, Dad, brother and sisters for their devotion and wholehearted

support.

Finally, to my beloved wife, Bin Wang, and our kids. I dedicate this thesis to them in their trust

and company.

Page 5: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

v

Contributions

1. A. Buin, P. Pietsch, J. Xu, O. Voznyy, A. H. Ip, R. Comin, E. H. Sargent. Materials

Processing Routes to Trap-Free Halide Perovskites. Nano Letters 2014, 14, 6281.

This work is mainly contained in Chapter 3. It focused on identifying the growth conditions

for producing low-trap-density, device-level films for planar PV, from theoretical and

experimental perspectives. A. Buin led the project and performed the DFT simulations. P.

Pietsch and I performed the experiments. I developed the material processing protocol of

using a new precursor to achieve the device-quality films and conducted the PL

measurements. I collaborated with Andrei to define the key questions and strategies in the

DFT studies.

2. J. Xu, A. Buin, A. H. Ip, W. Li, O. Voznyy, R. Comin, M. Yuan, S. Jeon, Z. Ning, J. J.

McDowell, P. Kanjanaboos, J.-P. Sun, X. Lan, L. N. Quan, D. H. Kim, I. G. Hill, P.

Maksymovych, E. H. Sargent. Perovskite-fullerene hybrid materials suppress hysteresis in

planar diodes. Nat Commun 2015, 6, DOI 10.1038/ncomms8081.

This work is mainly contained in Chapter 4. It focused on the materials chemistry of the

grain boundaries and electron-extraction interfaces to reveal the origins of hysteresis in

perovskite devices. I conceived the idea and conducted the experimental design, analysis, and

manuscript preparation, in collaboration with co-authors. Density functional theory

calculations were performed by A. Buin. Peter Maksymovych and Jon-Paul Sun carried out

cAFM studies.

3. J. Xu, O. Voznyy, R. Comin, X. Gong, G. Walters, M. Liu, P. Kanjanaboos, X. Lan, E. H.

Sargent. Crosslinked Remote-Doped Hole-Extracting Contacts Enhance Stability under

Accelerated Lifetime Testing in Perovskite Solar Cells. Advanced Materials 2016, 28, 2807.

This work is mainly contained in Chapter 5. It focused on the new design of a hole-

extraction interface to improve device stability and performance simultaneously. I initiated

the idea and conducted the experimental design, analysis, and manuscript preparation in

collaboration with co-authors. O. Voznyy, R. Comin, and X. Gong assisted the theoretical

studies and optoelectronic simulations.

Page 6: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

vi

Table of Contents

Table of Contents

Acknowledgments.......................................................................................................................... iv

Contributions....................................................................................................................................v

Table of Contents ........................................................................................................................... vi

List of Tables ................................................................................................................................. ix

List of Figures ..................................................................................................................................x

List of Common Acronyms and Symbols ..................................................................................... xii

Chapter 1 ..........................................................................................................................................1

Introduction .................................................................................................................................1

1.1 Solar photovoltaics: a carbon-neutral energy capture strategy ............................................1

1.2 From wafer-based to solution-processed photovoltaics .......................................................2

1.3 Theoretical and practical limits in photovoltaic efficiency ..................................................4

1.4 Materials interface engineering for solution-processed photovoltaics ................................6

1.4.1 Charge-separating interfaces ....................................................................................7

1.4.2 Interfaces within the absorbers: grain boundaries ...................................................8

1.5 Roadmap of thesis: engineering interfaces in perovskite photovoltaics ..............................9

Chapter 2 ........................................................................................................................................11

Perovskite photovoltaics ...........................................................................................................11

2.1 Crystal structure and charge transport ...............................................................................11

2.2 Device architectures ...........................................................................................................13

2.3 Challenges in perovskite photovoltaics .............................................................................14

2.3.1 Stability ..................................................................................................................14

2.3.2 Hysteresis ...............................................................................................................15

2.4 Research goals and methodology.......................................................................................15

Page 7: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

vii

Chapter 3 ........................................................................................................................................17

Materials processing for low-trap-density perovskite solids ....................................................17

3.1 Introduction ........................................................................................................................17

3.2 Perspective from DFT on perovskite growth .....................................................................17

3.3 Growth conditions control for low-trap perovskite film ....................................................22

3.3.1 Diffusion length .....................................................................................................23

3.3.2 Crystalline morphology .........................................................................................25

3.4 Conclusions ........................................................................................................................27

Chapter 4 ........................................................................................................................................28

Fullerene-perovskite interaction at electron-extraction interfaces ............................................28

4.1 Introduction ........................................................................................................................28

4.2 Improvement of hysteresis and photovoltaic performance ................................................29

4.3 Mechanistic studies of perovskite-PCBM interaction .......................................................31

4.3.1 Material characterization .......................................................................................31

4.3.2 DFT simulations.....................................................................................................32

4.4 Perovskite-PCBM mixture phase distribution ...................................................................33

4.5 Charge dynamics and hysteresis characterization ..............................................................35

4.6 Discussion: Ionic motion and hysteresis in perovskites ....................................................39

Chapter 5 ........................................................................................................................................40

Crosslinked hole-extraction interface improves hysteresis and stability ..................................40

5.1 Introduction ........................................................................................................................40

5.2 Crosslinked interface on perovskite top surface ................................................................41

5.3 Remote doping for hole-extraction conductivity ...............................................................44

5.4 Efficient PV with reduced hysteresis .................................................................................45

5.5 Mechanistic study of remote-doped hole-extraction ..........................................................48

5.5.1 Material characterization .......................................................................................48

Page 8: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

viii

5.5.2 Optoelectronic simulations ....................................................................................48

5.6 Improved stability under external stress ............................................................................50

5.7 Conclusions ........................................................................................................................54

Chapter 6 ........................................................................................................................................55

Conclusions ...............................................................................................................................55

6.1 Summary and Impact .........................................................................................................55

6.2 Outlook for perovskite solids and PV ................................................................................56

6.2.1 Maximum efficiency ..............................................................................................56

6.2.2 Long-term stability.................................................................................................57

References ......................................................................................................................................59

Appendices .....................................................................................................................................65

Page 9: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

ix

List of Tables

Table 3-1. Computed defect formation energies under (a) I-poor (Pb-rich) (b) I-rich (Pb-poor)

conditions. ..................................................................................................................................... 21

Table 4-1. Statistics of steady-state performance with different PCBM distribution and thickness

....................................................................................................................................................... 30

Page 10: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

x

List of Figures

Figure 1-1. Estimated potential of various renewable energy sources.. ......................................... 1

Figure 1-2. The Sun’s power spectrum reaching the Earth and the limits of solar conversion. ..... 4

Figure 1-3. Photogenerated charge carrier transport in a single-junction solar cell.. ..................... 5

Figure 1-4. The current–voltage characteristic and definition of important terms in photovoltaics

......................................................................................................................................................... 6

Figure 1-5. Architectures and interfaces in solution-processed solar cells. .................................... 8

Figure 1-6. Outline of thesis: material interfaces engineering in perovskite PV .......................... 10

Figure 2-1. Crystal structure of perovskites with the generic chemical formula ABX3.. ............. 12

Figure 2-2. Perovskite solar cell architectures evolution. ............................................................. 13

Figure 2-3. Instability and hysteresis in perovskite solar devices. ............................................... 14

Figure 2-4. The framework of full interfaces in the perovskite solar cells. ................................. 16

Figure 3-1. Surface states of the Pb halide perovskites. ............................................................... 18

Figure 3-2. Tetragonal perovskite, its formation from experimentally employed precursors, and

its defect energy levels. ................................................................................................................. 19

Figure 3-3. Formation energies and volume densities of key defects in tetragonal lead perovskites

....................................................................................................................................................... 22

Figure 3-4. Physical configuration of PbI neutral antisites in the tetragonal phase ...................... 23

Figure 3-5. Experimental investigation of transport in novel iodide-poor perovskite films. ....... 25

Figure 3-6. Non-continuous perovskite films grown using PbI2 precursor .................................. 26

Figure 3-7. Dense and smooth perovskite planar films grown using anhydrate lead acetate. ...... 26

Page 11: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

xi

Figure 4-1. Steady-state photovoltaic performance of an ultra-thin perovskite-PCBM hybrid film

....................................................................................................................................................... 29

Figure 4-2. Solution-processed planar device structures in this study .......................................... 31

Figure 4-3. Perovskite-PCBM hybrid process and in situ passivation mechanism.. .................... 32

Figure 4-4. 3D phase separation and homogeneous PCBM distribution in hybrid solid ............. 34

Figure 4-5. PCBM phase separation at perovskite grain boundaries ............................................ 35

Figure 4-6. cAFM study of hysteresis-ion relationship for control films and hybrid films.......... 36

Figure 4-7. Long-term steady-state dark current measurement of planar devices. ....................... 37

Figure 4-8. Effect of PCBM on charge carrier dynamics ............................................................. 38

Figure 5-1. Hole extraction contact employing material crosslinking and interface doping ........ 43

Figure 5-2. Thermally crosslinked VNPB is insoluble and enables the layer-by-layer deposition

....................................................................................................................................................... 44

Figure 5-3. Ultraviolet photoelectron spectroscopy (UPS) studies of VNPB layer ..................... 45

Figure 5-4. Improved photovoltaic performance with interface doping ....................................... 46

Figure 5-5. Electrical simulation of devices using interface doping ............................................ 49

Figure 5-6. Evolution of performance, morphology and material under external stress .............. 51

Figure 5-7. Assessment of device evolution under the external solvent attack ............................ 53

Figure 5-8. Evolution of material and morphology under the external solvent attack. ................ 53

Figure 6-1. Fraction of Shockley-Queisser detailed-balance limit for voltage and current

achieved by record cells. ............................................................................................................... 57

Page 12: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

xii

List of Common Acronyms and Symbols

PV – photovoltaic

PCE – power conversion efficiency

VOC – open circuit voltage

JSC – short circuit current density

FF – fill factor

MPP – maximum power output point

EQE – external quantum efficiency

DFT – density functional theory

Ef – Fermi level

Eg – semiconductor’s bandgap

WF – work function

VBM – valence band maximum

CBM – conduction band minimum

ETL – electron transport layer

HTL – hole transport layer

PCBM – phenyl-C61-butyric acid methyl ester

spiro-MeOTAD – 2,2’,7,7’-Tetrakis(N,N-di-p-methoxyphenylamine)-9,9’-spirobifluorene

VNPB – crosslinked N4,N4' -Di(naphthalen-1-yl)-N4,N4' -bis(4-vinylphenyl)biphenyl-4,4'-

diamine

DMF – N,N-Dimethylformamide

cAFM – conductive atomic force microscope

XPS – X-ray photoelectron spectroscopy

UPS – ultraviolet photoelectron spectroscopy

SIMS – secondary ion mass spectrometry

Page 13: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

1

Chapter 1

Introduction

1.1 Solar photovoltaics: a carbon-neutral energy capture strategy

Identifying clean energy sources to meet the world’s growing demand is one of society’s

foremost challenges. Projected worldwide population expansion and economic growth will more

than double the current global energy consumption rate to ~30 TW by 2050, even with

aggressive conservation. The threats of global warming and climate change due to excessive

greenhouse gas (such as CO2) emission impose a second requirement on new energy resources.

To balance the increased energy consumption rate and CO2 emission rate, 15 TW will need to be

derived entirely from zero-carbon-intensity (C-neutral) renewable energy sources by 20501.

Figure 1-1. Estimated potential of various renewable energy sources. For a fair comparison, the potential of each

power source (in units of W) is converted to “equivalent chemical fuel” power (in units of Wc). The estimated

efficiency of this conversion is included in the final comparison. In this regard, if the produced energy is already in

the form of a chemical fuel, such as in solar fuels, then there is no conversion. For energy in an electrical form, such

as in solar electricity, the conversion factor from the electrical to chemical form is 75 %. The conversion factor from

energies in a mechanical form, such as wind power, the conversion factor to a chemical form has a lumped efficiency

Ocean Thermal Gradient

Ocean Tidal

Ocean Salinity Gradient

Ocean SurfaceCurrent

2001 Supply

GeothermalHydropower

Ocean wave

Wind

SolarFuel

Solar Thermal

Solar Electricity

Sources with >15TW extractable and technical

potential

10-6 10-3 1 103 10610-6

10-3

1

103

106

Extractable Potential (TWc)

Tech

nic

al P

ote

nti

al (

TWc)

Page 14: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

2

of 25%. This conversion factor includes a 33 % efficiency associated with the conversion to an electrical form followed

by another 75 % efficiency associated with the conversion to the chemical form. Energy generated by heat sources,

such as from ocean thermal power, a Carnot efficiency ηc (second law of thermodynamics) is first used for a conversion

to a mechanical form, such as turbine rotation, followed by another 25 % efficiency associated with a conversion from

the mechanical form to chemical fuel, yielding a ηc·25 % lumped efficiency. Figure reproduced from ref. 1, copyright

2006, U.S. Department of Energy.

Solar energy can be directly captured from sunlight and converted into different forms such as

electricity, chemical fuels, and thermal energy. It dwarfs other renewable energy source

including hydro, ocean, wind, and geothermal. It has the greatest potential to meet the daunting

challenge of 15 TW C-neutral power (Figure 1-1). In fact, sunlight powers geological processes

(atmospheric motion, water and vapor transport, etc.) and therefore is largely responsible for

various secondary renewable energies (wind, ocean wave, and hydropower). In a theoretical

scenario, the total solar energy striking the Earth’s surface (~89 300 TW after the atmospheric

loss) within mere hours could meet the world’s energy needs for an entire year. A net 10%

efficient solar energy farm covering ~0.2% of Earth land would provide the global 15 TW C-

neutral power. Indeed, solar energy is the source with the technical potential safely exceeding 15

TW.

Solar electricity produced from solar photovoltaics is far from fulfilling its huge potential, and

the market is in a rapid growth phase today. The worldwide accumulated capacity of solar

photovoltaics was ~177 GW by 2014, with a total power output equal only to ~1% of the

worldwide electricity demand2.

Advances in science and technology provide ways to accelerate further cost-to-efficiency

reductions in solar photovoltaics. The use of photovoltaic power as an energy source also

requires breakthroughs in cost-effective and scalable energy storage and transport to match time-

and-space varying energy supply and demand3.

1.2 From wafer-based to solution-processed photovoltaics

Crystalline silicon (c-Si) solar cells dominate today’s PV production, having advanced greatly in

the past half-century. The best performance (over 25% in top lab-level devices) is based on the

highest-quality crystal wafers4,5. Because of the cost and stringent processes of making and

handling wafers, the production cost would be substantially reduced if devices could be obtained

Page 15: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

3

using films grown on glass or other inexpensive substrates. Silicon’s bandgap (1.1 eV) is indirect

and does not match the solar spectrum optimally, thus requiring a large thickness (~200 μm) of

material to completely absorb light with photon energies above the bandgap. Another wafer-

based solar cell, GaAs, can keep the absorber thickness relatively small (~2 μm) because of its

direct-bandgap character and its bandgap being close to the optimum (1.42 eV) and its high

absorption coefficient. GaAs solar cells hold the highest efficiency record (~28.8%) of single

junction PV cells4,6. The materials cost and energy-intensive processing (epitaxial growth using

chemical vapor deposition), however, so far restrict application to niche markets, such as space

technology.

Direct-bandgap, thin film devices represent second-generation photovoltaics. They benefit from

reduced materials usage and less stringent deposition processes. Because they use high

absorptivity direct-bandgap materials (polycrystalline CdTe, CuInxGa(1-x)Se2 or so-called CIGS,

amorphous Si), the thickness of the absorber can be reduced below one micrometer, ~100×

thinner than c-Si solar cells. The films can be deposited on glass, or flexible and lightweight

substrates, increasing possible applications (such as mobile and wearable) compared with rigid

wafer-based cells. Direct deposition is less energy-consumptive and is suitable to produce large-

area cells. Thin film cells are relatively lower in efficiency (~21% in best lab-scale device)

compared to c-Si cells7,8. The scarcity of elemental components (indium and gallium in CIGS

cells; tellurium in CdTe) may also impose limits on cost and scalability.

The development of solution-processable PV materials is a major research frontier in emerging

PVs, such as dye-sensitized solar cells (DSSC, or “Grätzel cell”)9–11, organic and polymer solar

cells (OPV) 12,13, colloidal quantum dot (CQD) photovoltaics14–16, and solution-processed bulk

inorganic PV such as copper zinc tin sulfide (CZTS) 17–19. Though their present efficiencies are

relatively low (~10-12%) and the stability of the absorber is often too short from a commercial

perspective, there is a lot of research invested in these technologies due to their potential as low-

cost, high-efficiency alternatives to present-day commercial modules. The goal of lower module

cost and energy cost are pursued by constructing devices on large, flexible substrates using low

temperature solution-processing. This enables economical and scalable manufacturing

techniques such as spray-painting, ink-jet printing, and roll-to-roll printing. Additionally,

flexible, lightweight modules may lower the installation and maintenance costs. The challenge

now is to continue progress in efficiency and increase stability.

Page 16: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

4

1.3 Theoretical and practical limits in photovoltaic efficiency

Achieving high efficiencies in photovoltaics requires an effective use of the broadband solar

spectrum (Figure 1-2)20. In the sun’s spectrum, nearly 50% of solar energy lies in the infrared

range. Natural photosynthesis uses part of the visible band, and its total solar energy conversion

efficiency is ~1%. Solar cells made using semiconductor junctions have the advantage of

broadband harvesting, spanning the visible and infrared, through bandgap engineering. A single-

junction solar cell made using a semiconductor with the optimal bandgap of ~1.34 eV has the

potential to reach a 33.7% solar-to-electricity conversion efficiency. This is known as the

Shockley-Queisser limit21.

Figure 1-2. The Sun’s power spectrum reaching the Earth and the limits of solar conversion. The AM1.5 solar

spectrum is broadband with distinct dips due to molecular absorption in the Earth’s atmosphere. The intensity of this

simulated solar spectrum is 1000 W m-2.20 Photosynthesis uses only the visible band, whereas solar cells can achieve

broadband absorption reaching out into the infrared region. Inset: photons with energy above the bandgap (Eg) are not

fully converted to electrical power due to the thermalization of charge carriers. The separation of quasi-Fermi levels

determines the open-circuit voltage Voc. Voltage loss relative to the bandgap Eg is caused by the spontaneous emission

in the requirement of thermodynamic detailed balance.

The Shockley-Queisser limit originates from a number of factors. Photons with energies below

the bandgap of the semiconductor are not absorbed, whereas photons of energy much above the

bandgap cannot be fully converted to electrical energy due to the thermalization of charge

400 800 1200 1600 2000 2400

1

2

Spectr

al pow

er

(W m

-2 n

m-1)

Wavelength (nm)

VB

CB

Eg qVoc

+

-

UV Visible Infrared

Page 17: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

5

carriers (Figure 1-2). Another physical limit is imposed by the thermodynamic detailed balance,

which requires the solar cell to be in equilibrium with its environment. There is spontaneous light

emission in a solar cell when it absorbs light. The corresponding radiative carrier recombination

forms a dark current that reduces the open-circuit voltage (VOC) to be well below the

semiconductor’s bandgap (Eg) (Figure 1-2 inset).

Multi-junction solar cells represent a strategy to overcome the thermalization limit present in

single-junction cells. Absorbers with different bandgaps are stacked sequentially to extract power

from their respective portions of photons in the solar spectrum, which then leads to higher

overall solar conversion efficiency. For example, a triple-junction cell could increase the

efficiency limit from 33.7% to 49%22. Epitaxially-grown triple-junction devices have reached

efficiencies of ~44% in the lab with the aid of concentrated sunlight23,24. Multijunction cells’

costs are high due to the complex structures and the high prices of materials.

In practical solar cells, more sources of loss reduce the efficiency. The achievable VOC is below

the Shockley-Queisser limit due to dark current contributions caused by carrier recombination in

the bulk and at interfaces in a solar cell (Figure 1-3)25,26. The short-circuit current, Jsc (see

definition in Figure 1-4) is also below the theoretical value: not all incident light is absorbed in

the active layer due to optical losses such as reflection and parasitic absorption. Furthermore, not

all generated carriers are collected due to non-idealities such as interfaces and contacts in a cell.

Scientific and technological advances are needed for efficiencies to approach more closely the

Shockley-Queisser limit.

Figure 1-3. Photogenerated charge carrier transport in a single-junction solar cell. Photoelectrons generated in

the p-type material on the left diffuse through the quasi-neutral region and then are swept by the built-in electric field

Efp

Ec

Ev

Efn

Sunlight

Electricity

Depletion regionQuasi-neutral Quasi-neutral

E

Electrode

Page 18: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

6

in the depletion region before they diffuse into the electron-collecting contact on the right. The dashed black arrow

represents charge carriers’ recombination pathway facilitated by trap states in the absorber and at the interfaces. The

corresponding dark current causes voltage and current losses in a practical solar cell. A key challenge is increasing

the diffusion length for minority carriers in solar cells. Ec and Ev indicate the band edges of the conduction band and

valence bands, respectively. Efp and Efn indicate the quasi-Fermi levels of holes and electrons respectively.

Figure 1-4. The current–voltage (J-V) characteristic and definition of important terms in photovoltaics.

1.4 Materials interface engineering for solution-processed photovoltaics

Advances toward increasing the efficiency of solution-processed solar cells rely largely on

controlling the abundant interfaces in devices, which, beyond the active layer, are primarily

composed of polycrystalline and amorphous layers25. Even though wafer-based crystalline Si PV

cells are considered to be planar and mature, the recent record performance (25.6%) was

achieved thanks to interface passivation of the crystalline Si (c-Si) surface using an ultrathin

layer of intrinsic hydrogenated amorphous Si (a-Si:H)5. In a classic solution-processed thin film

PV, multiple layers are stacked within a thickness of only a few hundred nanometers (nm), and

thus the conceptual difference between a “layer” and an “interface” is blurred. For clarity, the

interfaces in a device can be classified into two categories:

0

VOC

JSC

MPPJMPP

VMPP

Dark

Illuminated

Voltage

Cu

rren

t

Figures of merit in photovoltaics

Voc Open-circuit voltage. The voltage output by illuminated PV when its contacts are opened, i.e., there is no external load;

JSC The electron current flowing through the illuminated PV when its contacts are shorted;

MPP Maximum power point, where the product of current and voltage reaches its maximum.

FF Fill factor. The ratio of maximum power tothe product of VOC and JSC

PCE Power conversion efficiency. The ratio of power output at MPP to the solar power incident on PV device

EQE External quantum efficiency. The ratio of the number of electrons flowing through the PV per second under short-circuit condition to the number of photons illuminating the device each second

Page 19: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

7

a) Charge-separating interfaces: Interfaces between photo-conversion materials, i.e., the

junction between intimately contacted photoelectron donors and acceptors, if applicable;

and interfaces between photo-conversion layers and top and bottom contacts.

b) Interfaces within absorbers, i.e., grain boundaries between the crystalline domains in

the absorber.

Engineering of these interfaces, including the control of morphology, electronic and chemical

interactions, is applied to improve device efficiency through two aspects: a) enhance the

absorption of light trapped in photo-absorbers; and b) reduce the recombination of photo-

generated carriers when they transit through the sequence of interfaces before they reach the

contacts to the external circuit.

1.4.1 Charge-separating interfaces

Solution-processed PV materials are typically polycrystalline and show limited diffusion lengths

(5-500 nm) for the performance-limiting photocarrier – the minority charge carrier. This

thickness is not enough to achieve complete absorption of solar irradiation in a planar absorber.

To break this absorption-extraction compromise, nanostructured or mesostructured interfaces are

employed and have been demonstrated in research on DSSCs (Figure 1-5a). In a DSSC, only a

small fraction (~1%) of sunlight is absorbed if a monolayer of dye molecules is anchored on a

planar electrode. The dye-electrode contact area can be increased more than a thousand times by

using nanostructured electrodes made of high-bandgap metal oxide nanoparticles, such as

titanium oxide (TiO2), tin oxide (SnO2), and zinc oxide (ZnO). The absorption and extraction can

exceed 80% across the entire visible range9–11,27. This concept of large interface areas is also

pursued in organic and polymer solar cells, where donor and acceptor molecules are mixed to

form a bulk heterojunction (BHJ), a 3D interpenetrating network in the photoactive layer (Figure

1-5b), that results in a high quantum yield of exciton dissociation12,13.

Materials chemistry and the electronics of donor-acceptor interfaces are also important. In

DSSCs, for example, back recombination of the separated carriers across the large-area

nanostructured interface becomes more likely and results in low open-circuit voltage. Forming a

dense self-assembled monolayer (SAM) and infilling co-adsorbents at the oxide interface

passivates the exposed surface defects and impedes the backflow of charges efficiently28. This

leads to near-unity net collection of photo-generated carriers.

Page 20: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

8

Ultrathin (typically <10 nm) interlayers have been devised to improve the interface between the

photoactive layer and the contacts in organic and polymer PV. The photoactive layer’s surface

potential can be modified via an interface electrical dipole generated by the interaction between

the interlayer’s functional group and the photoactive layer29,30. The optimized energetic

alignment and built-in potential between the active layer and the contacts improves performance

by facilitating forward injection and suppressing back recombination.

Figure 1-5. Architectures and interfaces in solution-processed solar cells. (a) In dye-sensitized solar cells, a

monolayer of dye (orange dots) is anchored onto a 3D electrode such as mesoporous TiO2. (b) In organic solar cells,

the electron donor and acceptor materials are in intimate contact with each other, forming an interpenetrating bulk

heterojunction. (c) In colloidal quantum dot solar cells, the monodispersed nanoparticles (blue dots) are densely

packed to facilitate the charge carrier transport between quantum dots. (d) In inorganic bulk polycrystalline solar cells,

charge carriers diffuse through the grain boundaries in bulk absorbers to be extracted at top and down hetero-contacts.

For each cell, the panel on the left illustrates the device structure; the panel on the right illustrates the material energy

band diagram under open-circuit condition.

1.4.2 Interfaces within the absorbers: grain boundaries

Engineering the interfaces within the absorber is a large contributor to progress in colloidal

quantum dot (CQD) PV, where there exists grain boundaries throughout the film composed of

monodispersed nanocrystals (Figure 1-5c). Long organic ligands retain the colloidal stability of

quantum dots in solution, but isolate quantum dots from electronic communication with each

other. Replacing the long insulating ligands with short ligands (such as organic ligands with

TiO2

TiO2+Dye

Electrolyte

Pt

CQDs

AuTiO2

Au

ZnO

CdS

CIGS

TiO2

PCDTBT

PCBM

Al

a c

b d

Page 21: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

9

short anion- and thiol-end functional groups, or halide atomic ligands), a process termed ligand-

exchange, leads to denser packing of quantum dots without dot fusion, thus maintaining quantum

confinement, leading to remarkably improved carrier transport throughout the film. The short

ligand chosen also plays a crucial role in passivating trap states on quantum dot surfaces, thereby

providing a clean bandgap in quantum dot solid films that prolongs the lifetime of charge carriers

to be collected before they can recombine16,31.

Grain boundary passivation in CIGS or CZTS polycrystalline thin films (Figure 1-5d) is also

applied to improve performance32,33. This was implemented by controlling the band offset

between grain boundaries and the bulk. Materials processing such as elemental stoichiometry

tuning and post-annealing are also found to modify the grain structure, grain orientation, and

grain boundaries’ composition, providing a handle to reduce the surface trap density.

1.5 Roadmap of thesis: engineering interfaces in perovskite photovoltaics

Among emerging solar photovoltaics, solution-processed perovskite solar cells show particular

promise to fulfill the simultaneous goals of high efficiency and low cost. As introduced in

Chapter 2, organic-inorganic hybrid perovskite materials can be applied with simple solution-

processing at low temperature. Device efficiencies have increased from 3.8% in 2009 to ~18% in

early 201434,35, and progress shows no signs of abating. However, the materials and devices are

relatively new and far from being well understood. The problems of hysteretic behavior and

operation instability must be overcome in order to further push the efficiency toward its

theoretical limit while maintaining low cost. The photochemical and photophysical pathways of

these new materials need to be revealed. Advances will rely on understanding and managing the

abundant materials interfaces that make up these highly polycrystalline devices – the major focus

of this thesis (Figure 1-6).

An outline of the flow of this thesis is shown in Figure 1-6. The effects of material processing on

perovskite grain crystals and trap state formation are studied theoretically and experimentally

(Chapter 3) and provide guidance on precursor optimization and film processing to form device-

level films. The materials chemistry and resultant interface electronic engineering are then

applied to the grain boundaries, electron-extraction interfaces (Chapter 4) and hole-extraction

interfaces (Chapter 5) of perovskite films to improve the overall device performance

Page 22: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

10

progressively. Density functional theory (DFT) and nano-probing technologies were employed

for accurate simulation and direct observation of interface processes at atomic and nanometer

scales. This provided guidance to understand and optimize perovskite interfaces.

Figure 1-6. Outline of thesis: material interfaces engineering in perovskite PV.

Chapter 3Trap state reduction via

material processing

Chapter 4Electron-extraction

interfaces and grain boundaries

Chapter 5Hole-extraction

interfacesand device top surface

Materials interface engineering on planar perovskite solar cells to improve

hysteresis, stability and efficiency

Page 23: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

11

Chapter 2

Perovskite photovoltaics

The emerging field of perovskite solar cells has captured the attention of the photovoltaics

research community, with the record efficiency exceeding 18% by 2014. More importantly, this

high efficiency is achievable with materials produced from high-throughput and inexpensive

solution processes. With the potential to achieve even higher efficiency and lower cost,

perovskite solar cells are commercially attractive. However, the origin of the superb performance

is still not fully understood. The performance anomalies, such as hysteretic behavior and

unstable operation, need to be addressed in order to fulfill the theoretical maximum efficiency.

These challenges motivated me to research the abundant material interfaces that make up

perovskite photovoltaic devices.

2.1 Crystal structure and charge transport

The term “perovskite” refers to the class of compounds that have the same type of crystal

structure as CaTiO3. These materials have the general ABX3 stoichiometry, consisting of corner

sharing octahedral BX6 in three dimensions (3D), with cation A occupying the cuboctahedral

cavity in each unit cell (Figure 2-1). The structure provides many degrees of freedom for the

elemental composition as long as charge balance and atomic radius tolerance conditions are

satisfied36,37. The A sites function as “spacers” between the BX6 network and provide a freedom

to diversify the perovskite structure from 3D frameworks to 2D layers, 1D chains, and 0D

frameworks37–40. The perovskite structure and its distortions provide a materials platform with

finely tunable physical properties.

Page 24: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

12

Figure 2-1. Crystal structure of perovskites with the generic chemical formula ABX3. Organic or inorganic

cations occupy position A (blue) whereas metal cations and anions occupy the B (grey) and X (purple) positions,

respectively.

Within the family of perovskites, the organo-metal halide perovskites, especially the class of

methylammonium lead halide perovskites (MAPbX3, where MA indicates methylammonium,

CH3NH3; and X the halide, typically I, often with a small fraction of Cl or Br), are attractive as

solar energy harvesters due to efficient ambipolar transport, strong light absorption, ease of

solution-processed manufacturing, and widely tunable optical properties41–44.

Depending on the halide content, the bandgap of mixed halide perovskites (MAPb(I, Br, Cl)3)

can be tuned from ~1.6 eV (pure MAPbI3) to 3.2 eV (MAPbCl3). Smaller bandgaps can be

achieved using appropriate organic cations (e.g., formamidinium, CH(NH2)2) and inorganic

cations (e.g., Sn ). A smaller bandgap, toward 1.3 eV, is desirable because of the higher

Shockley-Queisser efficiency limit for single-junction devices.

The methylammonium lead halide perovskites have long diffusion lengths over 1 μm for both

holes and electrons, as measured in single crystal studies45,46. Although the quality of perovskite

polycrystalline films is highly dependent on the materials processing conditions, it is widely

agreed that a diffusion length much greater than 100 nm can be achieved via appropriate film

deposition47–49. Combined with a high absorption coefficient, the long diffusion length in

perovskites, dwarfing other solution-processed PV materials, enables a break of the “absorption-

extraction compromise” to construct highly efficient planar solar cells.

Page 25: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

13

Experimental reports also show that the photo-generated carriers in perovskites are primarily

present as free electrons and holes, rather than as excitons as in other solution-processed PV

materials (e.g., organic solar cells, colloidal quantum dot solar cells). The high dielectric constant

and low exciton binding energy are inherent in halide perovskites, and enable charge separation

at room temperature50,51.

Although not fully understood yet, these outstanding charge transport characteristics in solution-

processed perovskites, causing them to surpass most of their “low-cost” competitors, underpin

the success in PV applications.

2.2 Device architectures

Halide perovskite PVs were introduced in 2009 by Tsutomu Miyasaka34. Initial progress in

perovskite photovoltaics has benefitted from the pioneering works of Nam-Gyu Park, Michael

Graetzel and Henry Snaith on solid-state perovskite-sensitized solar cells in 201252,53.

Mesoporous scaffolds were used to minimize the limits imposed by minority carrier drift and

diffusion via inclusion of the active light absorber within nanometer-sized electron-harvesting

pores (Figure 2-2a). The structures were very similar to those of dye-sensitized solar cells

(DSSC). A worldwide effort to improve device performance with such architectures led to rapid

efficiency improvements in perovskite solar cells (from 3.8% in 2009 to ~18% in early 2014)

34,35,54, rapidly approaching the performance of commercial-grade silicon-based photovoltaic

modules.

Figure 2-2. Perovskite solar cell architectures evolution. (a) Cross-sectional structure of a solid-state perovskite-

sensitized solar cell where the thick mesoscopic metal oxide (TiO2 or Al2O3) scaffold is infiltrated by the perovskite.

(b) Cross-section of a planar heterojunction solar cell lacking the thick metal oxide mesoporous scaffold.

Electron transport contact

Transparent conductive oxide

Hole transport contact

Back contact

Glass substrate (front contact)

Mesoporous scaffold

Perovskite

a

Electron transport contact

Perovskite

Transparent conductive oxide

Hole transport contact

Back contact

Glass substrate (front contact)

b

Page 26: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

14

Planar-electrode (as distinct from mesoporous) devices (Figure 2-2b) have also attracted

extensive studies, based on the finding that perovskites act as efficient ambipolar charge-

conductors55,56. Compared with the mesoporous architecture, planar devices are particularly

important in certain realms of application, such as in photodetector arrays (which demand

stringent spatial uniformity pixel-to-pixel), lasers (which require planarity for minimized

scattering while waveguiding), and flexible photovoltaics (which strive to avoid high-

temperature mesoporous oxide processing)57,58.

2.3 Challenges in perovskite photovoltaics

2.3.1 Stability

A major problem in perovskites is the instability of the solar cell, both in materials and devices.

For the perovskite material, the organic cations are vulnerable to moisture and heating cycles,

and easily desorb from the lattice of the perovskite, leaving the inorganic framework of PbI2

(Figure 2-3a)59. It has also been suggested that the lead-halide bond is inherently not photo-

stable. In devices, UV-induced degradation is also linked to the redox photocatalysis effect at the

TiO2 surface and thus is more significant in mesoporous devices60,61. Such detrimental interfacial

interactions have also been shown at the perovskite-hole transfer interface62. The hole transfer

layer’s sensitivity to external variables is also an origin of device instability63,64.

Figure 2-3. Instability and hysteresis in perovskite solar devices. (a) The visible degradation from CH3NH3PbI3

perovskite (dark brown) to PbI2 (yellow) phase in air due to moisture. (b) Hysteresis of perovskite solar cells shown

in their J-V solar response.

Time in air

(a)

Voltage (V)

Cu

rren

t d

ensi

ty (

mA

cm

-2)

0.5 10.0

10

20

0

(b)

Page 27: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

15

2.3.2 Hysteresis

Severe hysteresis represents another big challenge for perovskites (Figure 2-3b). This is seen in a

scan-direction- and scan-speed-dependence to photo J-V characteristics65–67. This behavior

makes overestimating the solar-to-electricity efficiency likely. Moreover, the devices with

hysteresis also exhibit current degradation when they are operated under the steady-state

condition. The hysteresis is likely associated with charge recombination at defective interfaces

and grain boundaries48,68–72. The origins of hysteresis in devices are multifold and are far from

well resolved.

2.4 Research goals and methodology

My research vision is geared toward the exploration of novel materials and device designs to

develop hysteresis-free, stable, perovskite solar cells. Studies would include (1) origins of

hysteresis and instability and remedies thereto and (2) ways of making the materials and devices

more stable and hysteresis-free.

Intriguingly, the enhanced diffusion lengths in perovskite single crystals largely result from the

absence of structural and grain boundary defects that are present in perovskite polycrystalline

films73, suggestive of strategies of material processing and interface engineering to improve the

perovskite-film based PV devices.

Bearing this consideration in mind, I start with understanding the trap physics and formation

chemistry of perovskite polycrystalline films during materials processing via DFT simulation in

Chapter 3. Based on this knowledge, I further devise the materials engineering of the full set of

interfaces (grain boundaries, electron-extraction interfaces, and hole-extraction interfaces) in

perovskite devices (Figure 2-4) to improve the hysteresis and stability characteristics in order to

ultimately improve device efficiency (Chapter 4 - 7).

Page 28: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

16

Figure 2-4. The framework of full interfaces (green) in the perovskite solar cells. Full interfaces include the grain

boundaries within perovskite polycrystalline films and the electron- and hole-extraction interfaces bridging the

perovskite and the charge-accepting contacts.

Hole-accepting contact (Au/Ag, top reflecting contact)

Electron-accepting contact (ITO/FTO on glass, transparent)

Perovskite grains Grain

boundaries

Sunlight

Electron-extraction interface

Hole-extraction interface

Page 29: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

17

Chapter 3

Materials processing for low-trap-density perovskite solids

3.1 Introduction

Forming a low-trap-density film with good morphology is the starting point for making

perovskite PV devices and for further interfaces engineering. Recent reports have consistently

emphasized that the specific choice of growth conditions and chemical precursors is central to

achieving high-quality films. For example, MAPbI3 films made from a mixture of lead-chloride

(PbCl2) and methylammonium-iodide (MAI) exhibit more impressive diffusion lengths (1 µm)

than those made from single-halide approaches. However, the roles and mechanisms are poorly

understood. In this chapter, I use density functional theory (DFT) to explore the electronic levels

and formation energies of trap and defect states as the perovskite MAPbI3 film is grown. This

study provides a framework for understanding the impact of composition on the band-structure

of crystals incorporating vacancies, dopants, and interfaces. More importantly, it answers the

practical questions of identifying the dominant electronic traps formed during perovskite film

growth from solution, and how to choose the growth conditions and chemical precursors to form

PV films with reduced trap states.

3.2 Perspective from DFT on perovskite growth

DFT calculations were performed on tetragonal MA–PbI3 by using a supercell including 1728

atoms, which can accommodate various types of lattice and substitutional sparse defects. In line

with previous reports74–76, the electronic structure is confirmed to be direct at the Γ-point, with a

valence band maximum (VBM) state in an antibonding combination of I 5p (dominant character)

and Pb 6s orbitals (Pb 6s–I 5p, σ*), while the conduction band minimum (CBM) has mainly Pb

6p character.

The first part of the study aims at exploring the impact of the nanocrystalline morphology on the

band structure of MA–PbI3. In particular, one of the remarkable features of these materials is

that, in spite of their polycrystallinity, they have exceedingly sharp bandedges44. Electronic

Page 30: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

18

structure calculations on slabs of perovskite crystals were carried out to evaluate the impact of

surfaces on their band structure. The corresponding density of states (Figure 3-1a) shows no

states in the gap either for bulk or surface electronic structures. Indeed the highest-lying valence

band states in asymmetrically terminated stoichiometric slabs (Figure 3-1b) exhibit a high degree

of delocalization, as do the lowest-lying conduction band states (Figure 3-1c). A crucial insight

into the origin of this behavior comes from the computed surface energies: very low values of

∼10 meV Å-2 are obtained. The low surface energy indicates high stability, obviating the need

for a reconstruction of the (001) terminated surface and also avoiding the need to add adsorbates

(e.g., introduce new ligands) in order to remain inert. It was shown recently77 that the surface

energy of terminated (001) PbI2 slabs is indeed small and compares quite well to the value of

surface energy obtained in this study.

Figure 3-1. Surface states of the Pb halide perovskites. (a) The density of states for a 16-monolayer slab (surface)

compared to the case of bulk materials. The inset reveals that no electronic states emerge in the slab compared to the

bulk. The wave functions of surface states at (b) the valence band maximum (VBM) remain highly delocalized in the

case of 16-monolayer slabs, similarly to those at (c) the conduction band minimum (CBM) states. The highest

amplitudes of positive- (negative-) valued wave functions are indicated in yellow (blue). Please note that the CBM

and VBM states are separated due to asymmetric termination.

Looking at defects first required an estimation of the formation energy of the perovskite relative

to its decomposition into pure PbI2 and MAI phases (Figure 3-2a). The calculations show that the

formation energy of the perovskite relative to its decomposition into pure PbI2 and MAI phases

Page 31: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

19

has a low value of about −0.1 eV. This indicates that the material and its precursors are

energetically close to phase coexistence of MAI and PbI2 which is consistent with recent

experimental findings54,78 of a residual PbI2 phase remaining even after long annealing times. It

is also consistent with reports60,79–82 that the details of preparation conditions are important to

achieve the best-performing materials.

Figure 3-2. Tetragonal perovskite, its formation from experimentally employed precursors, and its defect

energy levels. (a) The energy of formation of the CH3NH3PbI3 perovskite (room temperature tetragonal structure is

shown) from its bulk precursors CH3NH3I and PbI2, with a kinetic barrier (transition state energy) also depicted.

Calculations discussed in the text report a −0.1 eV difference between the perovskite and its precursors, consistent

with experimental studies that show the presence of a secondary PbI2 phase. (b) Energy levels associated with the

defect states corresponding to neutral and charged vacancies (VPb, VI, VMA), neutral and charged interstitials (Pbi, Ii,

MAi), and neutral and charged states associated with antisites (PbI and IPb).

The electronic levels (Figure 3-2b) and formation energies (Hf) of various classes of defects were

then analyzed. The defects include vacancies (VPb, VMA, VI), interstitials (Pbi, MAi, Ii), and

antisites (PbI, IPb), where in the latter case AB indicates that A is substituted by B. The value

Page 32: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

20

of Hf was used to estimate the density of the relevant vacancy species, which will be proportional

to exp(−Hf/kT). One can see from Figure 3-3a and b that the major acceptor defects are VPb–2,

VMA−1 and Ii

–1, while donor defects are VI+1 and Pbi

+2, with an associated charge density

sufficient to induce doping anywhere from p- to n-type, depending on the chemical potentials.

Defects VI+1, VMA

–1, and MAi+1 possess the lowest formation energies over the entire bandgap.

The transition levels of VPb–1 and Ii

–1 are located on top of the VBM, indicating that negative

charge states are stable over the entire bandgap.

The defect formation energies of neutral defects are given in Table 3-1. One can also see that Pbi,

PbI, and IPb represent the “negative-U” defects83,84, indicating that these defects are not stable at

+1, −1, and +2 charge states, respectively. The large structural relaxation in the case of

PbI0 (interplanar I–I–I bridging) is consistent with such “negative-U” behavior of the charge

neutral PbI antisite. Figure 3-3a and b also show defect states with negative formation energies at

various Fermi levels. Usually, this implies that at these chemical potentials and Fermi levels it is

not possible to grow the crystal, and Fermi level pinning happens in such a way that all of the

defect energies are positive85 while satisfying the overall charge-neutrality condition; in other

words, the structure is not stable for such parameters86,87. However, these are values of the single

defect formation energies, and nothing prevents the complexation of single defects whenever the

latter is favorable in energy. Thus, defect clustering expands the range of chemical potentials and

Fermi levels, i.e., growth conditions.

So far, the complex defect formation has been reported in the case of VPbI2 alone88 and shown

recently to have low formation energies89. All vacancies produce either slightly perturbed states

in the bands (which do not capture carriers), or shallow traps and resonances (deep localized

states hybridized with conduction or valence band states90 within the band), implying that

carriers can still relax easily to VBM and CBM. In contrast, certain interstitials and antisites

associated with Pb and I form electronic states deep inside the bandgap. The most important

figure of merit is the density of deep traps in the semiconductor volume, which will determine

the rate of capture of charge carriers and of loss due to recombination. In semiconductors having

good electronic transport parameters, it is indeed the density of recombination centers that

directly controls the diffusion length of charge carriers91.

Page 33: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

21

Table 3-1. Computed defect formation energies under (a) I-poor (Pb-rich) (b) I-rich (Pb-poor) conditions.

Further analysis of the tetragonal vs. cubic crystal phase leads to a small number of qualitatively

and quantitatively different conclusions with respect to defect state energies. Overall the

agreement with previously published results79 is good, except that it was previously found that

the PbI neutral antisite defect has a high formation energy in the cubic case. In the tetragonal

phase, the PbI neutral antisite defects possess a much lower formation energy. Physically, this

difference arises because in the tetragonal phase two distinct PbI2 layers are found (Figure 3-4): a

nearly planar PbI2 plane and also a bulged PbI2 plane. The methylammonium cation has been

shown experimentally to exhibit a small reorientational barrier92 and to possess a small energetic

difference between two perovskite structures with different methylammonium cation

orientations, thus affecting the structure of the PbI2 layers.

The formation energy for charged defects, and consequently the volume density of the various

classes of localized electronic states, is a function of the Fermi level in the semiconductor. The

former is determined by whether the crystal is grown from iodine-poor or iodine-rich conditions,

which constitutes the central link between the growth chemistry and the charge transport

performance of these materials. In a crystal grown under I-rich conditions (Figure 3-3a), the

PbI antisite (Pb atom replaced by I) deep trap has a formation energy of less than 0.2 eV for all

choices of Fermi level between the valence and conduction bandedges, i.e., for all nondegenerate

doping conditions. As a prediction, a perovskite grown under I-rich conditions will show a high

density of deep traps that will curtail the diffusion length. In contrast, in a crystal grown under I-

Defect Ef, (eV) (a) Ef, (eV) (b)

2.69 0.47

0.87 2.04

0.85 2.05

1.91 0.62

1.12 2.43

1.05 3.27

1.75 0.57

3.6(4.22) 0.19 (0.9)

1.54 4.88

Page 34: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

22

poor conditions (Figure 3-3b), there exists a window of Fermi levels, EF ≥ 0.9 eV (measured

relative to the VBM), where all deep traps have formation energies that exceed 0.38 eV. This

places the equilibrium density of trap states below 1015 cm–3. At this volume density, traps are

spaced a median ∼200 lattice constants in all crystallographic directions, enabling diffusion

lengths above 100 nm. Such low trap densities are also consistent with the impressive open-

circuit voltages seen in the best-reported lead iodide perovskite devices.

Figure 3-3. Formation energies and volume densities of key defects in tetragonal lead perovskites. Defect

formation energies for iodine-poor (a) and iodine-rich (b) growth conditions. Continuous (dashed) lines denote

shallow (deep) traps (see legend). The red region in (a) indicates the range of Fermi energies where trap densities

exceed 1018 cm–3—showing that no Fermi level choice yields a semiconductor substantially free of deep traps in the

case of I-rich growth conditions. Conversely, in the case of I-poor growth (b), trap densities below 1015 cm–3 can be

achieved in the range of Fermi energies (green region). In the case of MA-related defects one has (a) MA-poor and

(b) MA-rich conditions.

3.3 Growth conditions control for low-trap perovskite film

From a chemical perspective, the presence in solution of simple ions (Pb2+ and I–) combined with

that of lead-iodide complex anions such as PbI3–,93 PbI4

2–, and PbI53– produces a motif similar to

the PbI0 neutral antisite defect which corresponds to bridging between three iodine anions in-

plane (Figure 3-4b) and interplane (Figure 3-4c) in the perovskite lattice. The motif is produced

indirectly, by violating local stoichiometry. The I–I–I angle is 172°, and the I···I bond length

between two iodines is 2.98 Å. Knowledge of this geometry allowed me to look for signatures of

this complex, as did the signature wave function pattern (I 5p–I 5p, σ*). The concentration of

PbI3– in solution is higher for PbI2 + MAI than for PbCl2 + MAI. One interesting direction these

Page 35: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

23

results suggest is that alternative Pb non-iodine-containing precursors, such as Pb(SCN)2,

Pb(CH3CH2)4, and Pb(C2H3O2)2 (dehydrated lead acetate, Pb(Ac)2), can provide another avenue

to reaching the I-poor conditions required for high-quality perovskites.

Figure 3-4. Physical configuration of PbI neutral antisites in the tetragonal phase. (a) Two different PbI2 planes

in the tetragonal phase. One of them is bulged. Upon structural relaxation of the PbI0 defect, two conformations are

possible: (b) in-plane bridging or (c) interplanar bridging. The bridge configuration has a motif very similar to the

I3– triiodide complex, with an I–I–I angle of 172° and bond length 2.98 Å. This allowed association of the PbI

0 neutral

antisite with macro-ion complexes such as PbI3– and PbI4

2– in solution-phase precursors, and these are expected to be

present in growth under iodide-rich conditions.

3.3.1 Diffusion length

As a way to validate this prediction, and in order to challenge experimentally the proposed

theoretical findings, I investigated the diffusion length in perovskite films formed using a

dehydrated lead acetate precursor (see methods in Appendix 1.1-1.3) in place of the typical

PbI2 or PbCl2. The use of this precursor satisfies the requirement for I-poor conditions while

providing a Cl-free platform in order to rule out any possible effect from residual amounts of

incorporated chlorine. Similarly to the mixed-halide case, I find that acetate does not incorporate

Page 36: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

24

into the final product. In order to probe the diffusion lengths for MA–PbI3 grown from Pb(Ac)2, I

excited the perovskite film using a 442 nm laser while measuring the photoluminescence (PL)

signal in a reporter layer of small bandgap quantum dots (see inset of Figure 3-5a). The quantum

dot PL intensity is directly controlled by the carrier diffusion length and by the thickness of the

perovskite film94.

Qualitatively these diffusion length experiments can be understood as follows: for thin

perovskite films with thickness d smaller than the carrier diffusion length LD, the amplitude of

the PL signal correlates positively with LD, because an increase in layer thickness results in a

higher absorption and hence a higher exciton generation rate, and at the same time almost all

excitons can diffuse through the perovskite and reach the reporter layer. For increasing thickness,

the generation rate reaches saturation because the absorption cannot exceed 100%, but fewer

excitons can now diffuse to the reporter layer. The PL profiles for different perovskite film

thicknesses are presented in Figure 3-5a. The overall behavior of PL intensity vs. film thickness

(Figure 3-5b) can be modeled using the following function:

𝑷𝑳(𝒅; 𝒂, 𝑳𝑫) ∝𝑳𝑫𝟐

𝟏−(𝒂𝑳𝑫)𝟐[𝒂𝒆−𝒂𝒅 +

(𝒆−𝒂𝒅

𝑳𝑫) 𝐬𝐢𝐧𝐡(

𝒅

𝑳𝑫)−𝒂

𝐜𝐨𝐬𝐡(𝒅/𝑳𝑫)] (3-1)

Here LD is extracted by fitting the experimental data of Figure 3-5b once the independently

measured absorption coefficient α is known (see inset). The diffusion length of the

Pb(Ac)2 based perovskite films was determined to be ~600 nm, thus considerably larger than

∼200 nm (even lower values of ∼100 nm have been reported in literature47) for films formed

using the PbI2 precursor. This suggests that chlorine itself does not play a central role in large

diffusion lengths, whereas the reduction of iodine content during the film formation process is

likely key to the remarkable charge transport in these materials.

Page 37: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

25

Figure 3-5. Experimental investigation of transport in novel iodide-poor perovskite films. (a) Measured

photoluminescence (PL) spectra for various perovskite layer thicknesses as indicated in the legend in the case of

Pb(Ac)2 precursor. The inset shows a schematic of the diffusion length measurement method: the sample is illuminated

at a wavelength (442 nm) that is strongly absorbed in the perovskite layer; photogenerated charge carriers diffuse to

the quantum dot reporter layer where they recombine, providing a spectrally distinct signature of their arrival. (b)

Experimental plot (black and red circles) of the corrected PL peak amplitude vs. perovskite layer thickness both for

the PbI2 and Pb(Ac)2 precursors. A fit based on Equation (3-1) allows estimating the diffusion length to be LD ≈ 600

nm and LD ≈ 200 nm, which is comparable to prior pure and mixed-halide films. Inset: measured absorption coefficient

of the perovskite absorber, where the excitation wavelength is highlighted.

3.3.2 Crystalline morphology

Recent reports have consistently emphasized that in planar devices without a mesoporous metal

oxide scaffold, it is challenging to form continuous and smooth perovskite films via solution

processing41,42,54,78,95. The resultant shunting pathways facilitate carrier recombination and limit

the device performance. Our observations in films grown directly from regular PbI2 protocols are

aligned with these reports. A non-continuous film is usually obtained that exhibits a high

percentage of porosity (Figure 3-6).

Page 38: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

26

Figure 3-6. Non-continuous perovskite films grown using PbI2 precursor. (a) Top view of a rough perovskite film

with a high percentage of porosity. (b) and (c) Cross-sectional images show grains in films with different thickness.

The films grown from the new protocol using anhydrate lead acetate (see methods in Appendix

1.1-1.3) exhibited different crystalline behavior. The resultant films were dense and smooth

(Figure 3-7). The film roughness was smaller than 10 nm when the film thickness was bigger

than 200 nm (inset of Figure 3-7a). The pin-hole-free characteristic was maintained even when

the film thickness was reduced down to ~100 nm (Figure 3-7c). Smooth and dense films

spanning a large range of thicknesses are desirable for planar optoelectronics, such as solar cells

and light emitting diodes (LED).

Figure 3-7. Dense and smooth perovskite planar films grown using anhydrate lead acetate. (a) A dense and

smooth perovskite film with thickness > 200 nm and roughness ~8 nm as characterized by AFM (inset). (b) The

zoomed in image of the film morphology. (c) The cross-sectional SEM of a continuous planar film with thickness

~100 nm, deposited on a rough FTO substrate.

Page 39: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

27

This work shows that by controlling the perovskite growth conditions, such as via anion

engineering, we can tune the crystalline kinetics, film formation, and therefore device

performance. My work of using anhydrate lead acetate (Pb(C2H3O2)2 ) to resolve the morphology

problem in planar films was supplemented by parallel efforts by my peers, where hydrate lead

acetate (Pb(CH3COO)2·3H2O) was used96.

This study clarified fundamental aspects underlying the impact of growth conditions on the

performance of MA–PbI3 films. It revealed delocalization of the electronic states within the local

nanocrystal surfaces that preserved the integrity of the bulk bandgap. The DFT-based analysis of

defect formation energies provided an explanation for the observation of a larger charge

diffusion length in perovskites prepared using iodide-free precursors78,97–99compared to MAI +

PbI2 growth conditions.

3.4 Conclusions

In summary, iodine-rich growth conditions were found to be prone to forming perovskites with a

high density of deep trap states (recombination centers). The use of mix-halide precursors help to

reduce the formation of key defects (Pb atom substituted by I). I-poor conditions were

constructed using an anhydrate lead acetate (Pb(C2H3O2)2) precursor and showed films with

better diffusion lengths and morphologies than for the single-halide method which is I-rich,

under the same deposition conditions. This is an important aspect when considering the control

over growth conditions for making efficient perovskite devices.

Page 40: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

28

Chapter 4

Fullerene-perovskite interaction at electron-extraction interfaces

4.1 Introduction

In Chapter 3, I identified via DFT simulation that I-rich conditions might be a dominant source

of deep trap states during perovskite growth from precursor solutions. I found that using lead

acetate (Pb(Ac)2) as the lead precursor would help construct I-poor conditions and form films

with lower traps and longer diffusion lengths. I engineered films based on these findings and

constructed purely planar perovskite PV devices. I focused on planar devices for two reasons.

First, there is no fundamental limit on perovskites for building planar devices without

mesoporous scaffolds considering the ambipolar transport and large values of diffusion lengths

in the perovskite material. Planar architectures also help minimize surface recombination and

increase the open-circuit voltage. Second, solution-processed planar perovskite devices are

highly desirable in a wide variety of optoelectronic applications such as in photodetector arrays

(which require stringent spatial uniformity pixel-to-pixel), lasers (which require planarity for

low-scatter waveguiding), and flexible photovoltaics (which strive to avoid high-temperature

mesoporous oxide processing)55–58,95,100.

I found, however, perovskite films on planar devices are prone to hysteresis and current

instabilities, which is highly consistent with recent reports in literature (details see Chapter

2.3.2). The performance of planar perovskite devices has, widely agreed to date, suffered from

two potentially interrelated concerns: severe hysteresis41–43; and, relatedly, recombination, likely

associated with defective grain boundaries induced by excess halides48,68–70,101. A dependence of

hysteresis on device architectures has also been observed, where inverted structures have

typically shown less serious hysteresis than regular planar devices, but have a lower open-circuit

voltage71,72.There has been, to date, no consensus as to the origins of these findings.

In this chapter, I pursue a solution-phase in situ passivation strategy with the goal of enabling

simple low-temperature materials processing and efficient passivation throughout the grain

boundaries in the perovskite active layer and electron-extraction interfaces. I also employ density

Page 41: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

29

functional theory (DFT) and nano-probing techniques to reveal the underlying sources of

hysteresis.

4.2 Improvement of hysteresis and photovoltaic performance

In the course of device studies of mixed materials made from solutions containing both

perovskites and the electron-acceptor PCBM (a derivative of C60, phenyl-C61-butyric acid

methyl ester), I observed an enhancement in photovoltaic performance (Figure 4-1a-c) and a

reduction in hysteresis (Figure 4-1d-e) relative to control devices based on perovskites alone, and

also compared to separate-layer PCBM-perovskite devices (Table 4-1, see Figure 4-2 for

different device structures). To create the mixed-material films, I dispersed PCBM and lead

acetate (Pb(Ac)2)48 in various ratios and formed films using a one-step deposition

process53,100,102,103 employing MAI as the organohalide precursor. As well as observing reduced

hysteresis (Figure 4-1d-e), I observed in the perovskite-PCBM mixed-material device a

substantial voltage enhancement (~ 0.1 V) (Figure 4-1a) and a higher fill factor compared to the

PCBM-free and bilayer PCBM-perovskite controls (see methods in Appendix 1.1-1.3).

Figure 4-1. Steady-state photovoltaic performance of an ultra-thin perovskite-PCBM hybrid film. (a) The steady

state open circuit voltage, VOC, (b) steady state short circuit current density, JSC, and (c) the steady state power

0.0 0.2 0.4 0.6 0.8 1.0 1.2-10

0

10

20Control

FF(Forward)=0.66

FF(Reverse)=0.42

Me

asu

red

cu

rre

nt

de

nsity (

mA

cm

-2)

Voltage (V)

0.0 0.2 0.4 0.6 0.8 1.0 1.2-10

0

10

20Hybrid

FF (Forward)=0.74

FF (Reverse)=0.70

Me

asu

red

cu

rre

nt

de

nsity (

mA

/cm

2)

Voltage (V)

(d) (e)

Maximum Power Point

0 20 404

6

8

10

12

14

16

Hybrid

Control

Ste

ad

y-s

tate

Eff

icie

ncy,

PC

E(%

)

Time, t(s)

(a) (b) (c)

0 5 10 15

0.9

1.0

1.1

1.2

Hybrid

Control

Op

en

Cir

cu

it V

olta

ge

, V

OC

(V)

Time, t(s)0 2 4 6

12

13

14

15

16

Hybrid

Control

Sh

ort

Cir

cu

it C

urr

en

t, J

SC

(mA

cm

-2)

Time, t(s)

Absorber thickness:150±10 nm

500nm

400 600 800

0

20

40

60

80

100

Measure

d E

QE

(%

)

Wavelength (nm)

Calculated Jsc=15.4mA cm-2

Page 42: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

30

conversion efficiency, PCE, of the perovskite-PCBM hybrid film (red) compared with the control perovskite only film

(blue). During steady-state measurement, the integrating time for each point is 0.35 second. (d) The instantaneous J-

V curve of the control device (perovskite film) with high hysteresis. The thicker curve indicates forward scan starting

from open circuit conditions; thin curve is the reverse scan from short circuit conditions. The scanning rate is 0.2 V s-

1. The fill factor (FF) of the forward scan is 66% while the reverse FF is reduced to 42%. The black point indicates

the ‘maximum power output point (MPP)’ is measured from the steady state PCE as shown in (c). The MPP here is

located between two instantaneous J-V curves due to the significant hysteresis and current decay. (e) The J-V scan of

a hybrid device shows very low hysteresis and low current loss, as shown in (b). The FF for forward (reverse) scan is

74% (70%). The steady-state MPP is consistent with the forward J-V curve, which demonstrates the stability of the

hybrid film. The inset of figure (e) shows the external quantum efficiency (EQE) of a hybrid device. The current

density predicted from the EQE is 15.4 mA cm-2, consistent with the steady state current density measured in (b). The

inset figure of (d) shows the thickness of the active layer in both devices is around 150 nm.

Table 4-1. Statistics of steady-state performance with different PCBM distribution and thickness.

Device configuration Steady-state performance Instantaneous

Type Thickness (nm)

VOC (V)

JSC (mA cm

-2)

PCE (%)

FF (%) Forward/Reverse

Control

150±20

0.97±0.02 13.1±0.8 6.7±0.5 62/38±3

Bilayer 1.08±0.02 14.2±0.4 10.6±0.4 71/64±3

Hybrid 1.09±0.02 14.4±0.4 10.9±0.4 72/65±2

Hybrid champion 1.11 14.6 11.9 73/68

Control 300±20

0.98±0.02 14.4±0.8 8.1±0.5 65/40±3

Bilayer 1.06±0.02 16.1±0.4 12.0±0.5 72/56±3

Hybrid 1.07±0.02 17.3±0.4 13.6±0.6 73/66±3

Hybrid champion 1.086 18.0 14.4 75/69

Statistics for each case are based on 20 devices prepared on separate substrates.

Page 43: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

31

Figure 4-2. Solution-processed planar device structures in this study. (a) Control device with PCBM-free pure

perovskite (CH3NH3PbI3) as the active layer; TiO2 and Spiro-OMeTAD as the electron transport layer (ETL) and the

hole transport layer (HTL), respectively.(b) Perovskite-PCBM bilayer structure with PCBM cast on TiO2 before

perovskite deposition; and (c) Perovskite-PCBM hybrid device with mixed material as the active layer.

4.3 Mechanistic studies of perovskite-PCBM interaction

4.3.1 Material characterization

I proceeded to seek mechanistic insights regarding the role, or roles, of the PCBM. Specifically, I

asked whether the PCBM could interact with certain chemical species in the mixed-material

solution; and whether studies of incorporation into films using the new process indicated a

homogeneous distribution of PCBM throughout the active layer, compared to segregation into a

bilayer device with PCBM either substantially below or above the perovskite.

Solution-phase spectroscopy provides one means to study the formation of complexes of PCBM

with the various perovskite solution-phase precursors. When PCBM is mixed into the normal

perovskite precursor solution, the bright yellow solution (Figure 4-3a left) turns to dark brown

(Figure 4-3a right). The absorption spectrum of perovskite-PCBM hybrid solution shows a peak

at 1020 nm (Figure 4-3b). This is in contrast with pure PCBM in the same solvent, which is

observed to be transparent in this wavelength region. The 1020 nm spectral feature is associated

in literature reports with the formation of a PCBM halide radical (Figure 4-3b inset)104–106.

FTO/Glasss

Perovskite-PCBMHybrid solid

Spiro-OMeTAD

FTO/Glasss

TiO2

Perovskite

Spiro-OMeTAD

Au

FTO/Glasss

Perovskite

Spiro-OMeTAD

PCBM

Au

TiO2 TiO2

Au

(b) Perovskite-PCBM Bilayer

(a) Control PCBM-free

(c) Perovskite-PCBM Hybrid

Page 44: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

32

4.3.2 DFT simulations

This reconfirmation of the strong PCBM-iodide interaction motivated me to explore, using

density functional theory (DFT), what might occur in a solid material. I looked in particular at

reactions PCBM might participate in at the excess-halide-associated defects at grain boundaries

previously reported to be a dominant source of electronic traps in lead methylammonium iodide

perovskites48,68,69,101. I focused specifically on the Pb-I antisite defect, in which iodine occupies

the Pb site and forms a trimer with neighbouring iodine atoms (Figure 4-3c)48. DFT reveals that,

with the introduction of PCBM near such Pb-I antisite defects, the wavefunction of the ground

state (Figure 4-3d) is hybridized between the PCBM and the perovskite surface. The bonding of

PCBM to defective halides is thermodynamically favoured and that this suppresses the formation

of deep traps (Figure 4-3e).

Figure 4-3. Perovskite-PCBM hybrid process and in situ passivation mechanism. (a) Pristine perovskite solution

(left) comprised of Pb(Ac)2 and MAI in dimethylformamide (DMF) solvent is bright yellow; the formulated

perovskite-PCBM hybrid solution (right) is brown; Simple one-step spin-coating is used to convert the hybrid solution

DO

S (

a.u

.)

Ev Ec

Trap state

600 800 1000 1200 1400 1600

0.0

0.5

1.0

No

rma

lize

d a

bso

rba

nce

(a

.u.)

Wavelength (nm)

PCBM in normal solvent

PCBM in Hybrid solution

Ac-e-

I- MA+

Pb2+

(a)

(b)

(c)

(d)

(e)

Grain boundary

Pb-I antistite

Pb I

Page 45: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

33

into an hybrid solid film, and the perovskite is in situ passivated by PCBM during self-assembly; (b) UV-Visible

absorption spectroscopy of the hybrid solution shows the interaction between PCBM and perovskite ions. PCBM

radical anion’s absorption peak at 1020 nm is identified in hybrid solution (red); while PCBM in same solvent (black)

has no absorption peak in this wavelength region; Inset of (b) shows details of such interaction: In hybrid solution,

electron-transfer is induced between the perovskite anions (I-) and PCBM and will result in PCBM radical anions and

PCBM-halide radicals. (c) A schematic of in situ passivation of halide-induced deep traps: PCBM adsorbs on the Pb-

I antisite defective grain boundary during perovskite self-assembly. (d) The wavefunction overlap shows the

hybridization between PCBM and the defective surface, enabling the electron/hole transfer for absorbance and

passivation. (e) DFT calculation of the density of states (DOS) shows that deep trap states (black) induced by Pb-I

antisite defects are reduced and become much shallower (red) upon the adsorption of PCBM on defective halides. Ec,

minimum of conduction band; Ev, maximum of valence band.

4.4 Perovskite-PCBM mixture phase distribution

Next I sought to determine whether the PCBM is distributed throughout the entire thickness of

the active layer that had been formed from the mixed perovskite-PCBM solution (Figure 4-4a).

Secondary ion mass spectrometry (SIMS) was used to probe the depth profile of PCBM and

perovskite. Pb and Ti are used as indicators of the perovskite and of the TiO2 substrate,

respectively. Since PCBM does not contain elements to identify it uniquely, I used instead for

this portion of the study a thiophene-containing derivative, [60]ThCBM ([6,6]-(2-Thienyl)-C61-

butyric acid methyl ester), which permits the use of sulfur as the tracer element107. The

[60]ThCBM is present homogeneously throughout the thickness of the hybrid film, with a

uniform concentration as a function of depth (Figure 4-4b). Using X-ray diffraction (XRD), I

found that the perovskite lattice diffraction peaks of the hybrid film are consistent with that of

control films without PCBM (Figure 4-4c). In addition, the average perovskite grain size in

hybrid films, estimated from the XRD peak width, is comparable to that of control films.

Also with the nature of the mixed material in mind, I employed Kelvin probe studies to examine

the work function (WF) of mixed-material films. The work function of the mixed-material films

lies between that of the pure perovskite and pure PCBM. Its value varies monotonically along

this continuum as a function of PCBM fraction incorporated. When very high PCBM fractions

are employed, evidence of phase separation and impacts on film morphology emerge: the PCBM

phase aggregates at perovskite grain boundaries and becomes clearly evident (Figure 4-5).

Page 46: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

34

Figure 4-4. 3D phase separation and homogeneous PCBM distribution in hybrid solid. (a) Scheme of planar

perovskite solar cell using perovskite-PCBM hybrid solid as the active absorber; PCBM phase is homogeneously

distributed at grain boundaries throughout the perovskite layer. (b) Secondary Ion Mass Spectrometry (SIMS) depth

profile of perovskite-PCBM hybrid film on TiO2 substrate showing homogeneous distribution of PCBM throughout

the film. The sputter etching begins at the air/film interface. PCBM is tracked by S element using analogous

[60]ThPCBM; perovskite is tracked by Pb element; TiO2 is tracked by Ti atom. (c) XRD patterns of the pristine hybrid

solid film (red) and the control perovskite film without PCBM (black). TiO2 compact layer on FTO is used as substrate.

XRD shows that in the hybrid solid, the perovskite crystal lattice is the same as the control film, and thus PCBM only

exists at the grain boundaries and interfaces throughout the film. (d) The transient photoluminescence of the hybrid

film. Pumping from the top of the film (black) and pumping from the bottom of the film (red) give identical signals,

showing homogeneous PCBM distribution. The hybrid film displays dense grains and full-coverage as observed via

SEM (inset left); the surface is ultra-flat with roughness ~6 nm as characterized by AFM (inset right).

These findings prompted me to posit the following picture of the mixed material. Perovskite

grains are formed with similar size and crystallinity with and without the PCBM (XRD). The

PCBM is distributed uniformly throughout the thickness of the film (SIMS), presumably in

0 100 200 300

0.01

0.1

1

No

rma

lize

d P

L (

a.u

.)

Time (ns)

Pump from Top

Pump from Bottom

(a)

0 100 200 300 400

0

1

No

rma

lize

d S

IMS

Co

un

ts (

a.u

.)

Sputter Time (a.u.)

Pb

S

Ti

TiO2 PCBM

Perovskite

(b)

10 20 30 40 50 60

XR

D in

tensity (

a.u

.)

2theta (deg)

Control:Perovskite

Hybrid:Perovskite:PCBM

(c)

(d)

PCBM

Perovskitegrains

TiO2 (ETL)

Spiro (HTL)

1um 1um

RMS=6.5nm

Page 47: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

35

between the grains. The PCBM could bind iodide-rich defects sites on these grain boundaries

(DFT), and/or could simply bind up excess iodide from the solution. From an electronic

standpoint, the incorporation of the PCBM throughout the film influences its work function in

proportion with the PCBM-perovskite ratio (Kelvin Probe).

To seek further indications regarding the extent of electronic interactions between the PCBM and

the perovskite grains in the mixed material, I acquired transient photoluminescence for pure

perovskites, PCBM-perovskite bilayers, and mixed materials, investigating each for the case of

excitation from each side. The mixed-material film shows identical transient PL traces for both

top and bottom illumination schemes (Figure 4-4d). In contrast, perovskite films with PCBM on

one side only exhibit different PL lifetimes when pumped from the different sides. The

invariance of the PL lifetime with top/bottom-side photoexcitation for the mixed-material system

agrees with SIMS, and further indicates that the hybrid film behaves as a homogenous

optoelectronic material throughout its thickness.

Figure 4-5. PCBM phase separation at perovskite grain boundaries. (a) Bulk phase separation is evident when

very high PCBM fractions are employed. (b) Zoomed in view of a region where PCBM phase emerges at grain

boundaries.

4.5 Charge dynamics and hysteresis characterization

A series of conductive AFM studies provide added spatial resolution of the electronic properties

of the films under study. I carried out the cAFM studies under high vacuum and dark conditions

to rule out the effect of light and moisture. By overlapping the grain topography and the

electrical current map of the films (Figure 4-6a and Figure 4-6e), I find that conductivity is

greatest at grain boundaries, both in the pure-perovskite and in the mixed-material films.

30um

PCBM phase

(a) (b)

Page 48: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

36

However, the mixed-material films have much higher conductivity near grain boundaries at

positive bias voltages, consistent with the evidence of the electron-transport medium PCBM

accumulating near grain boundaries and providing continuous pathways for electron egress. I

also obtained I-V traces at various spatial positions, and find that control perovskite films exhibit

major hysteresis behavior when scanned in the reverse bias direction (Figure 4-6b, 4c and 4d).

Given the pure perovskites’ slow response on the seconds timescale, the hysteretic I-V curves are

consistent with the proposed hysteresis mechanism of ionic transport in perovskite solids108–111.

To my knowledge, this is the first direct experimental observation of memristive properties

within the perovskite material itself via cAFM111. This observation is generally in agreement

with the very recently reported ionic motion processes in CH3NH3PbI3 perovskite materials71,112.

In contrast, in perovskite-PCBM mixed films, the hysteresis effect is greatly suppressed under all

conditions (Figure 4-6f, 4g and 4h). These observations further substantiate a picture in which

PCBM influences electronic properties when it associates with the perovskite grains at their

grain boundaries.

Figure 4-6. cAFM study of hysteresis-ion relationship for control films and hybrid films. (a) and (e) the gray-

scaled contact-mode AFM (background) with overlaid color-scaled conductive AFM images (positive sample bias

voltage: 1 V). (b-d) I-V hysteresis of control film increases when increase the negative bias and injected current (solid

line: forward sweep, dashed line: reverse sweep). (f-h) I-V hysteresis of hybrid film is suppressed when increasing the

negative bias and injected current. Scanning rate is ~0.5 V s-1 (see Methods).

Page 49: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

37

Additional device studies offer further information about the role of PCBM in perovskite device

performance and hysteresis. Planar devices incorporating PCBM – whether at an interface or

throughout in bulk – are consistently superior in performance to control devices without PCBM

(Table 4-1). Incorporating the PCBM into the film becomes even more advantageous to collected

current for thicker active layers, suggesting that the PCBM accepts photocharges and assists in

their extraction to the TiO2. The champion planar devices were obtained using the perovskite-

PCBM mixed material and exhibited steady-state PCE exceeding 14.4%, 1.5 times more efficient

than my PCBM-free perovskite controls.

I also investigated the reverse saturation current density in the various devices and found that the

hybrid films consistently reduced the dark current by 2 orders of magnitude (Figure 4-7).

Rectifying behavior is also maintained much longer in the mixed material compared to

perovskite controls.

Figure 4-7. Long-term steady-state dark current measurement of planar devices. Perovskite-PCBM hybrid

devices (red and black) and control devices (blue and cyan) are tested under reverse bias -0.5 V. Hybrid devices

showed almost 2 orders of magnitude lower dark current and no breakdown during the course of the measurement.

Bias is applied continuously and dark current is sampled every 1 second.

0 5000 10000 15000 20000

1E-9

1E-8

1E-7

1E-6

1E-5

1E-4

Da

rk C

urr

en

t (A

)

Time (S)

Hybrid

Hybrid-2

Control

Control-2

TiO2 Peorvskite Spiro

Page 50: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

38

These last observations motivate further evaluation of the role of PCBM in trap passivation at

perovskite grain boundaries. I used transient photovoltage to quantify the prevalence of mid-gap

trap states in each class of materials and devices (details see Methods). I obtained a notably

longer carrier lifetime over a wide range of photovoltages in the mixed material (Figure 4-8a).

This indicates reduced non-geminate recombination for the perovskite-PCBM hybrid films. I

also compare the transient photoluminescence of hybrid films to investigate the impact of PCBM

on carrier extraction. When the PCBM-perovskite hybrid ratio is increased progressively, the PL

exhibits consistently greater quenching, indicating efficient electronic coupling between the

well-dispersed PCBM phase and the perovskite (Figure 4-8b, orange, pink, and red curve).

When the PCBM ratio is extremely high (Figure 4-8b, black curve), the photoluminescence

quenching efficiency began to degrade greatly. Significant phase segregation occurs, with the

appearance of large PCBM domains that lack effective interconnectivity for carrier extraction. I

concluded that comprehensive incorporation of PCBM in the interstitial volumes among grains

in the perovskite system is required (i.e. sufficient PCBM material miscibility in the perovskite

solid is needed) to produce continuous pathways for carrier extraction to enhance

performance113.

Figure 4-8. Effect of PCBM on charge carrier dynamics. (a) Charge carrier lifetime of hybrid device (red) and

control device (blue), determined from transient photovoltage measurement under open-circuit condition. (b)

Transient photoluminescence of hybrid films with increasing PCBM ratio progressively (orange, pink, red, black)

compared with control film on glass (blue), showing the enhanced electron extraction. The quenching efficiency

increases monotonically with increasing hybrid ratio, indicating the increasing PCBM-perovskite interfaces. When

continuously increasing the PCBM hybrid ratio (black), the quenching efficiency begins to reduce abnormally, due to

0.6 0.7 0.8 0.9 1.0 1.1

1E-5

1E-4

1E-3

Hybrid

Control

Re

com

bin

atio

n lifetim

e (

S)

VOC (V)

(a)

0 100 200

1E-3

0.01

0.1

1

No

rmalized P

L (

a.u

.)

Time (ns)

Hybrid (1:200)Hybrid (1:100)Hybrid (1:50)

Hybrid (1:10)Control

(b)

Page 51: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

39

the emergence of large domains of agglomerated PCBM, reducing the effective interconnectivity between perovskite

and PCBM.

4.6 Discussion: Ionic motion and hysteresis in perovskites

I close with a discussion of mechanisms likely at work, and one more speculative mechanism, in

the mixed-material films. My data suggests that PCBM, when incorporated at or near perovskite

grain boundaries, makes a significant impact on electronic properties. The transient

photovoltage, combined with the DFT analysis and the spectroscopy showing PCBM radical

formation, suggest that PCBM plays a passivating role at iodide-rich trap sites on the surfaces of

these grains. At the same time, the long timescale of hysteresis in pure perovskite films and its

substantial suppression in the mixed material, combined with the vastly lower reverse dark

current in the mixed material, suggest an additional effect at work in addition to the passivating

role. I propose that ions, such as the iodide anion, can potentially migrate under an applied

electric field, producing an ionic current. This can explain the slow response of hysteresis108,111

and the instability of the dark current when pure-perovskite and bilayer devices are employed.

By tying up iodide-rich surface sites, or simply unincorporated iodide anions, PCBM can reduce

anion migration through defects at grain boundaries108,109. This rearrangement under external,

and also built-in internal, electric fields, could account for solar cell hysteresis. For example,

when the device is poised at the JSC condition, the large built-in field may induce anionic charge

motion that works against this field, leading to a drop in photocurrent in time. A relatively rapid

scan towards VOC will therefore suffer from low photocurrents; whereas, following an extended

pause at VOC, during which anions can diffuse back to equilibrium positions, a rapid scan to JSC

will feature a high current in view of the lack of charge compensating the built-in field.

Page 52: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

40

Chapter 5

Crosslinked hole-extraction interface improves hysteresis and stability

5.1 Introduction

So far, I identified that ionic motion may be a major source of hysteresis and instability in

perovskite devices, and this field-induced ionic motion was also found to be highly related with

the defects at interfaces and grain boundaries (Chapter 4). Given the major strides in active layer

engineering and hole-extraction interfaces, there exists now the opportunity to improve hole-

extraction interfaces to enable further progress in hysteresis and stability improvement.

The top surface (the hole-extraction interface in the device) of the perovskite is of particular

interest in the long-term stability of devices, because it is exposed directly to external stresses.

Specifically, since hybrid halide perovskites have a highly ionic character, they can decompose

under external stresses such as moisture, solvents, and heating cycles, especially if not fully

encapsulated41–43. The resultant ionic complexes are then highly reactive with transition metal

oxides114 (such as MoO3) and metal contacts115,116. Materials engineering strategies, such as

adding crosslinking among perovskite grains117, have been shown to enhance the stability of the

active material, including in the presence of moisture.

In perovskite devices employing a top hole-extracting contact, the engineering of the HTL (hole

transport layer) offers an opportunity to add protection to the perovskite that underlies it59,62,118–

120. The HTL should desirably be robust to external stresses, such as high operating temperatures,

and at the same time should efficiently facilitate hole extraction and thus promote overall device

performance. It should also be transparent to produce rear-metal-contact reflections and also to

enable semi-transparent and multi-absorber devices121,122.

The organic HTL widely used in many top-performing perovskite solar cells, spiro-MeOTAD

(2,2’,7,7’-Tetrakis(N,N-di-p-methoxyphenylamine)-9,9’-spirobifluorene), requires the ionic

dopant Li-TFSI (Bis(trifluoromethane)sulfonimide lithium salt) with additive tBP (4-tert-butyl

pyridine)35,53,54,78,103,117,123–126. This additive has been found to evaporate at 85°C59,63,64, limiting

devices’ thermal stability and also curtailing their capacity to withstand subsequent processing

Page 53: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

41

steps. Further, this doping mechanism, which involves interactions with oxygen, requires fine

control63,127–129. The additive tBP and ionic dopant Li-TFSI have been found to interact with the

ionic perovskite layer and contribute thereby to the undesired introduction of water into the

active layer, thus contributing to perovskite device degradation59,62,129. Small pinholes in spiro-

MeOTAD layers were recently identified116,120,129 and found to facilitate the migration of iodine-

containing compounds from the perovskite, leading to corrosion of the metal top contact. Finally,

the sensitivity of spiro-MeOTAD to solvents imposes severe constraints on subsequent solution-

phase processing steps atop the device121.

To go beyond reliance on sensitive Li-doped spiro-MeOTAD atop the perovskite, alternatives

such as opaque carbon-based hole-extracting contacts119, hybrid carbon nanotube-polymers59,

and inorganic CuSCN130,131 have been investigated. These have shown improved device stability

and chemical robustness. To date, however, the benefits of these alternatives have come with

costs to performance: they have each quantitatively degraded solar cells’ open-circuit-voltage,

hysteresis, and fill factor. For example, the fill factors shown in these devices59,61,119,131 are

typically appreciably below the benchmark value (~75%) achievable in the state-of-art devices

employing doped spiro-MeOTAD35,54,103,117,123,124,126. These compromises to performance have

been ascribed to poor band level alignment and inefficient egress of charges across the resultant

interface.

In this Chapter, I pursued a new HTL strategy with the goals of protecting the perovskite,

achieving the needed free carrier density and work function without the use of chemical dopant

additives, and ultimately achieving high-performance perovskite solar cells that would exhibit

enhanced stability.

5.2 Crosslinked interface on perovskite top surface

My approach (Figure 5-1a) employed crosslinking of the polymer HTL in contact with the

perovskite in order to render the material insoluble and thermally stable. I would achieve the

needed deep work function and high hole free carrier density via a remote doping strategy (see

detailed methods in Appendix 1.5).

I focused the HTL work on arylamine derivatives, for these feature a HOMO (Highest Occupied

Molecular Orbital) level similar to the ionization potential of the perovskite (5.4~5.5eV). I first

Page 54: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

42

explored UV-crosslinkable arylamine derivatives based on cationic ring-opening polymerization

of oxetane groups132–135, for these are known to be substantially inert when they are in intimate

contact with underlying active materials. The crosslinking process employs a cationic

photoinitiator to break the C-O bonds within each oxetane group under UV radiation. It thereby

constructs an insoluble network by forming a crosslinking C-O bond between different oxetane

groups.

Unfortunately, I witnessed much lower photovoltaic performance when compared to devices that

used conventional Spiro-MeOTAD. I proposed that known by-products132,133 produced in situ by

the cationic photo-acid initiator used in the UV-crosslinking process degrades the electronic

quality of the perovskite. I proposed that the organometal halide perovskites are particularly

sensitive and thus not immune to the cationic photoinitiator.

The observed in situ degradation motivated me to devise a perovskite-compatible crosslinking

agent. I focused on non-ionic polymerizable groups and found that by thermally inducing

crosslinking between styrene groups136, I could form crosslinked films using the new arylamine

derivative (N4,N4' -Di(naphthalen-1-yl)-N4,N4' -bis(4-vinylphenyl)biphenyl-4,4'-diamine),

which I term VNPB.

When VNPB was deposited using spin-casting, it allowed me to form crosslinked films that did

not evolve by-products that would degrade the underlying perovskite (Figure 5-1b). The needed

crosslinking proceeds under mild thermal conditions and does not require the use of an initiator:

instead, crosslinking is achieved by an addition reaction through the opening of the double bonds

in styrene groups of adjacent VNPB units. Multiple styrene groups in each VNPB unit enable the

formation of a three-dimensional network with good coverage and strength. Compared to the

corresponding NPB film that lacks crosslinking groups, the VNPB film is insoluble and

thermally stable, factors that enable the stacking of VNPB films via a layer-by-layer solution

process (Figure 5-2). Thermally-induced polymerization of styrene groups has also recently been

used to enhance an organic electrode interlayer that resides under the perovskite. Consistent with

my findings, the crosslinked interlayer showed remarkable resistance to solvent stress

(perovskite-soluble polar solvents such as DMF) and annealing-stress while casting the

perovskite atop137.

Page 55: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

43

Figure 5-1. Hole extraction contact employing material crosslinking and interface doping. (a) Two-step scheme

to form the insoluble and thermally-stable hole extraction contact. In the first step, the organic hole transport layer

(HTL) is deposited and then thermally crosslinked; in the second step, an interface doping layer is simply deposited

atop of the HTL and doping is achieved via the interface charge transfer. (b) Details of the thermal crosslinking

process: double bonds (red lines) in styrene groups in the hole transport layer (VNPB) are opened and then crosslinked

via an addition reaction, thereby forming an insoluble, thermally-stable film. (c) Schematic of interface doping:

ground-state electron transfer occurs from the hole transport layer, having low ionization-energy, to the interface with

the high electron-affinity material, in this case transition metal oxide MoO3, thereby enhancing the hole carrier density

throughout the thin HTL. (d) Device structure of the planar perovskite device using a VNPB-MoO3 double-layer as

the top hole extraction contact. (e) The SEM cross-sectional image shows the full device covered by a dense hole

extraction layer based on the VNPB-MoO3 double-layer stack. (f) High resolution TEM further resolves the fine

interface of the VNPB-MoO3 double-layer. VNPB and MoO3 are confirmed to be in a dense and amorphous phase,

forming a smooth interface with the underlying polycrystalline perovskite layer.

Interface electron transfer

a

b c

[ ]

N N

[ ]

n n

N N

… … ……

Thermal crosslink

VNPB

d

FTO on Glass

TiO2/PCBM

Perovskite

VNPBMoO3

Metal (Au)

200nm

e

10nm

Perovskite

MoO3

VNPB

Au

Interface doping layer

(MoO3)

Ener

gy +

-

Hole transport

layer

Ef

Ef

f

Page 56: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

44

Figure 5-2. Thermally crosslinked VNPB is insoluble and enables the layer-by-layer deposition. (a) Cross-

sectional SEM image of 2 layers of crosslinked VNPB and (b) 4 layers of crosslinked VNPB film deposited by spin-

casting in a layer-by-layer fashion. It confirms that the final film exhibits a bulk morphology, and there is no detectable

interface between adjacent layers. Noted that for ease of observation, the concentration of VNPB used in this test is

much higher than that used in device hole transport layer.

5.3 Remote doping for hole-extraction conductivity

The VNPB crosslinked layer is intrinsic, and thus incapable of efficient hole extraction from the

perovskite. I sought to introduce free holes, and to do so without chemical doping used in Spiro-

MeOTAD59,62,116,120,129. I pursued an interface remote doping strategy138, wherein I deposited a

deep-workfunction transition metal oxide layer atop the HTL (Figure 5-1a). The dense and

chemically-inert crosslinked VNPB would serve to keep physically separate, and thereby prevent

chemical reactions among, the perovskite and the metal oxide (MoO3) layers114. The free hole

density would be introduced into the otherwise-intrinsic HTL via ground-state electron transfer

to the deep-workfunction metal oxide (MoO3 in this work) at the organic-inorganic interface

(Figure 5-1c)139,140.

High resolution microscopy confirmed intimate contact between the VNPB and MoO3 (Figure

5-1d-f), a precondition for efficient interface electron-transfer. The VNPB-MoO3 double-layer

structure was also confirmed to be in a dense amorphous phase and thus suffered no issues of

lattice mismatch with the polycrystalline perovskite. Ultraviolet photoelectron spectroscopy

studies (UPS) confirmed that the HOMO level of the VNPB layer (-5.4~-5.5 eV) is highly

aligned with perovskite (Figure 5-3), enabling a substantially barrierless hole extraction pathway.

Page 57: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

45

Figure 5-3. Ultraviolet photoelectron spectroscopy (UPS) studies of VNPB layer. UPS was carried out using He I

(21.22 eV) photon lines from a discharge lamp. The applied bias is 15 eV. The VNPB film is tested on an Au substrate.

The statistics analysis of multiple samples gives the HOMO level = -5.46 ± 0.08 eV.

5.4 Efficient PV with reduced hysteresis

The robust crosslinked hole transport layer (VNPB), coupled with an inorganic metal oxide layer

(MoO3) in a double-layer fashion, not only provides thermally-stable and solvent-resistant

protection for the perovskite, but also provides a stable and efficient doping process that leads to

a 16.5% solar PCE measured at steady state (Figure 5-4).

In planar devices that employed the new contact strategy (see methods in Appendix 1.5), I

observed highly stable steady-state photovoltaic performance when the devices were operated at

their maximum power point (Figure 5-4c, red curve; see test details in Appendix 2.1). Their

performance was equivalent to the benchmark devices that employed spiro-MeOTAD hole

transport layers (Figure 5-4c, blue curve). External quantum efficiency (EQE) measurements

were carried out and agreed with the measured current densities.

20 15 10 5 0

I. (

a.u

.)

Binding energy (eV)

3 2 1 0 -1

Binding energy (eV)

He I (21.22eV)

Page 58: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

46

Figure 5-4. Improved photovoltaic performance with interface doping. (a) The instantaneous J-V curve of control

devices using doped Spiro-MeOTAD (blue) as the hole transport layer, compared with devices using undoped Spiro-

MeOTAD (grey). The undoped Spiro-MeOTAD leads to a sharply reduced fill factor (FF) and performance. Arrows

indicate the voltage scanning direction. The thicker curve is the forward scan starting from open circuit condition

while the thin curve is the reverse scan starting from short circuit condition. The scanning rate is 0.2 V s-1. The inset

of (a) illustrates Spiro-MeOTAD doping by the use of Li salts (blue dot) throughout the film. (b) The J-V curve of

newly-designed devices using the VNPB-MoO3 interface doping hole-extraction contact (red), compared with devices

using VNPB alone (grey). Without the interface doping, the FF and overall performance show a stark decline. In

contrast, the device using interface doping shows an 80% FF with negligible hysteresis. The inset of (b) illustrates the

doping at the interface (red region) of the VNPB-MoO3 double-layer. (c) Steady state power conversion efficiency

(PCE) operated at the maximum power point of devices using interface doped VNPB (red squares), doped Spiro-

MeOTAD (blue circles), undoped VNPB (grey triangles) and undoped Spiro-MeOTAD (grey diamonds). The stable

current output at maximum power point indicates no hysteresis, which is consistent with the observation in the test of

instantaneous JV shown in (b). The decay of steady-state PCE of undoped devices is typically associated with

hysteresis in JV curves. (d) UV-Visible-IR absorption spectroscopy of the VNPB-MoO3 double-layer (red) showing

the signature of the interface charge-transfer-complex (CTC) in the near-infrared absorption region, while MoO3

(black) and VNPB individual layers (grey) exhibit no absorption features in the same wavelength region. Inset of (d)

illustrates that the CTC (red region) resides at the interface when VNPB is covered by the interface doping layer

MoO3. (e) The photoluminescence (PL) quenching effect in a VNPB-MoO3 double-layer (red) versus a VNPB single

0 50 100 150

4

5

6

7

8

10

15

20

Doped-Spiro

Doped-VNPB

Undoped-Spiro

Undoped-VNPB

Time, t(s)0.0 0.2 0.4 0.6 0.8 1.0 1.2

-10

0

10

20

Cu

rre

nt

den

sity (

mA

cm

-2)

Voltage (V)

Li-saltdopant

Spiro

Perovskite

TiO2

Control

Doped

Undoped

0.0 0.2 0.4 0.6 0.8 1.0 1.2

-10

0

10

20

VNPB

MoO3 Interface doping

Perovskite

TiO2

Crosslinked

Doped

Undoped

Steady-state efficiency, PCE(%)c

e

a b

d

400 800 1200 1600 2000

0

10

20

Ab

so

rpta

nce

(%

)

Wavelength (nm)

MoO3

VNPB

VNPB-MoO3

Interface charge-transfer-complex (CTC)

MoO3

VNPB

0 5 10 15 20 25 30

10-2

10-1

100

VNPB

VNPB-MoO3

Tra

nsie

nt P

L c

ou

nt (a

.u.)

Time (ns)

400 500 600

PL

coun

t (a.

u.)

(nm)

Page 59: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

47

layer (grey), induced by the interface charge-transfer-complex, is observed in transient and steady-state PL

measurements (inset) of VNPB. The arrow indicates the VNPB PL peak (450 nm) where transient PL was measured.

I observed that the instantaneous JV-curve provides a > 80% fill factor with negligible hysteresis

in planar devices (Figure 5-4b, red curve). From a statistical analysis on a large sampling of

devices that featured the new interfacial remote-doping hole-extraction contact, the average

hysteresis was found to be low (2%), an improvement from my control devices that had used

spiro-MeOTAD (6%). The low hysteresis agrees with the concept that the inert perovskite-

crosslinked interface reduces chemical actions such as those seen with chemically-doped spiro-

MeOTAD59,62,116,120,129. Here I quantify the hysteresis using Equation (5-1):

𝐇𝐲𝐬𝐭𝐞𝐫𝐞𝐬𝐢𝐬 = (𝐀𝐫𝐞𝐚𝒇𝒐𝒓𝒘𝒂𝒓𝒅

𝐀𝐫𝐞𝐚𝒓𝒆𝒗𝒆𝒓𝒔𝒆− 𝟏) × 𝟏𝟎𝟎% (5-1)

where Area𝑓𝑜𝑟𝑤𝑎𝑟𝑑 (Area𝑟𝑒𝑣𝑒𝑟𝑠𝑒) is the integrated area under the forward (reverse) scanning JV

curve. Low hysteresis correlates with stable steady-state output power. Devices with high

hysteresis in JV curves show decay of output current and power when operated at steady-

state65,118,141. To avoid any overestimation of PCE arising from JV hysteresis, I report the solar-

to-electricity efficiency only using the steady-state power-to-power performance of devices

operated at their maximum power point under constant AM1.5 solar illumination.

The high fill factor and low hysteresis indicate efficient charge extraction in the best-designed

remote-doped devices. In contrast, the fill factor was seriously compromised in all undoped

controls (Figure 5-4a and 2b, grey curves). The steady-state performance of undoped controls is

low, decays further during testing, and is accompanied by high hysteresis (Figure 5-4c). The

same trend of device degradation occurs in both classes of undoped devices (spiro-MeOTAD

devices without Li salt mixing; and crosslinked VNPB devices without a MoO3 interface layer).

This reconfirms the crucial role played by the MoO3 layer in introducing across-the-interface

remote doping in the crosslinked VNPB layer.

Page 60: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

48

5.5 Mechanistic study of remote-doped hole-extraction

5.5.1 Material characterization

I sought mechanistic insights into the role of the interface doping. Specifically, I investigated the

physical picture of interface doping at the VNPB-MoO3 organic-inorganic heterojunction, and

explored the design criteria for this new double-layer hole extraction structure.

UV-Vis-IR spectroscopy provides one means to study electron transfer at the interface: when

MoO3 is deposited on top of a VNPB film, I observe a near-infrared absorption peak that I

associate with the interface charge-transfer-complex142–145. This is in contrast with either

individual VNPB or MoO3 layers, which on their own are transparent in this wavelength region

(Figure 5-4d). This sub-bandgap absorption feature has previously been associated in literature

reports with the formation of intermediate states induced by charge-transfer-complexes at the

donor-acceptor interface. The attribution of this spectral feature to ground-state charge-transfer-

complexes at the interface is further verified by photoluminescence (PL) quenching effects

observed both in steady-state PL (Figure 2e, inset) and time-resolved PL measurements (Figure

5-4e). The PL tests reveal that the interfacial charge-transfer-complexes behave as quenching

sites for excitons in VNPB films through polaron-exciton quenching146–148. Consistently, in

spiro-MeOTAD, sub-bandgap absorption, PL quenching and time-resolved-PL quenching are

observed only when spiro-MeOTAD is doped using the Li-salt (see doping method in Appendix

1.6). In VNPB-MoO3, charge-transfer doping is accomplished at the interface alone, and

therefore the absolute parasitic absorption is much less than that in bulk doped spiro-MeOTAD.

These findings further agree with the picture of interface doping via ground-state electron-

transfer at the VNPB-MoO3 heterojunction depicted in Figure 5-1c.

5.5.2 Optoelectronic simulations

Next I sought insights into the role of interface doping in devices. I made use of self-consistent

optoelectronic device simulations149 and looked particularly at the organic-inorganic interface of

VNPB-MoO3 in the steady state. The band-bending of VNPB and MoO3 layers (Figure 5-5a and

3b) reconfirms the p-type doping of VNPB via interface electron-transfer. The theoretically-

predicted J-V behavior (Figure 5-5c) of solar devices with and without a MoO3 interface doping

layer are in excellent quantitative agreement with experimental results (Figure 5-4b). The

Page 61: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

49

simulations also predict that the interface doping layer will increase photovoltaic performance,

most significantly through the fill factor of devices. When I tuned the HOMO level of the hole

transport layer (Figure 5-5d), I observed linear control over the doping effect: fill factor increases

when the HOMO of the hole transport layer moves toward the HOMO of the perovskite layer.

VNPB satisfies this design rule very well, as confirmed by UPS measurements of the HOMO

level (Figure 5-3). With respect to the choice of work function, there exists a wide performance-

insensitive region (Figure 5-5e and 3f), requiring only that the workfunction of the interface

doping layer be deeper than the HOMO of the HTL (~-5.4 eV). The hole extraction efficiency,

associated with fill factor, declines sharply only when the workfunction of the doping layer is too

shallow to accept electrons. Fortunately, the interface doping material MoO3 resides - even when

one accounts for the spread in reported workfunctions58,139,140,146,147,150- in the performance-

insensitive region (-5.4 ~ -7 eV).

Figure 5-5. Electrical simulation of devices using interface doping. (a) The equilibrium-state energy band diagram

of devices using interface doping hole extraction contacts (VNPB-MoO3). Ec (Ev) indicates the edge of the conduction

(valence) band while Ef and red line denote the Fermi level. (b) Expanded view of the band alignment and band

0.0 0.5 1.0

0

10

20

Interface

doping layer

workfunction =4.6eV

5.2

4.85.0

>5.4eV

Cu

rre

nt D

en

sity (

mA

cm

-2)

Voltage (V)

-5.0 -5.2 -5.4 -5.6

60

80

FF

PCE

Hole transport layer

HOMO level (eV)

FF

(%)

12

16

20

PC

E (%

)

HOMO of VNPB

-4

-3

-2

-1

0

1

2

3

4

5

Ec

Ev

EFE (

eV)

0.0 0.2 0.4 0.6 0.8 1.0 1.20

5

10

15

20

Cu

rre

nt D

en

sity (

mA

cm

-2)

Voltage (V)

VNPB

VNPB-MoO3Ec

Ev

EF

VNPB

MoO3

-5 -6 -7

20

40

60

80

FF

PCE

Interface doping layer

workfunction (eV)

FF

(%)

10

15

20

PC

E (%

)

HOMO of HTL

HOMO of perovskite

a b c

fd e

TiO2

Perovskite

VNPB/MoO3

Au

Page 62: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

50

bending at the interface of the VNPB-MoO3 double-layer stack. (c) Performance comparison between devices with

(red) and without (grey) interface doping layers. Without interface doping, the fill factor remarkably decreases,

consistent with experimental observations (Figure 2b). (d) Performance evolution when the HOMO level of the hole

transport layer (HTL) changes. The fill factor increases when the HTL HOMO is well aligned with the HOMO of

perovskite (dash line). The arrow indicates that the HOMO level of crosslinked VNPB (5.46 ± 0.08 eV measured from

UPS, Figure S3) is highly aligned with the HOMO level of perovskite, and therefore is expected to result in the

optimized performance when used as the HTL. (e) Performance dependence on the workfunction of the interface

doping layer. The fill factor is unaltered in a rather extended range (-5.4 ~ 7 eV), as long as the workfunction of the

interface doping layer is deeper than the HOMO of the hole transport layer (dash line). (f) J-V curves corresponding

to the performance evolution shown in (e) showing that the performance drop occurring when the workfunction of the

interface doping layer becomes shallower than the HTL HOMO. The reported MoO3 workfunction range resides in

the optimal performance region.

5.6 Improved stability under external stress

I then proceeded to assess the enhanced stability of the new devices under external stresses such

as heating, moisture, and solvent. Devices with a remote-doped crosslinked top contact (VNPB-

MoO3) were first investigated under thermal stress. Most striking is the devices’ retention of

their superior performance (maintenance of at least 95% of initial performance) and low

hysteresis behavior following fully 1 hour of annealing at ~100˚C (Figure 5-6b). Under the same

stress, control devices with conventionally-doped spiro-MeOTAD lost more than 30% of their

performance irreversibly. This came principally through a severe degradation in fill factor and an

increase in hysteretic behavior (Figure 5-6a). Consistently, the rectification, under dark

conditions, of conventional devices also degraded irreversibly. Similar trends as those for

hysteresis degradation and performance decay in control devices were also observed in long-term

steady-state performance testing. Just as in the undoped-spiro-MeOTAD devices (Figure 5-4a),

the serious degradation in fill factor and the hysteresis indicates the loss of doping efficiency

under such thermal stress. Additional direct evidence came from an optical microscopy study: I

observed a new crystalline pattern in doped spiro-MeOTAD films (Figure 5-6c) following the

thermal stress test. The irreversible morphology degradation is linked to the phase separation of

dopant and spiro-MeOTAD host, a change that coincides with outgassing of the tBP additive at

temperatures that exceed 85˚C59,63,64. In contrast, the morphology of the VNPB-MoO3 double-

layer is thermally stable (Figure 5-6d), attributed to the robustness of both the crosslinked

material and the interface remote-doping mechanism.

Page 63: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

51

Figure 5-6. Evolution of performance, morphology and material under external stress. (a) The performance of

devices using Spiro-MeOTAD as the hole-extraction contact tested at room temperature (grey) and after a 110 ˚C

burn-in test (blue) [In the burn-in test, devices are annealed at 110 ˚C for 1 hour in an N2 environment and tested after

cooling down to room temperature]. (b) The performance of devices using VNPB-MoO3 tested at room temperature

(grey) and after 110 ˚C burn-in process (red). (c) Optical microscopy (reflection mode) of a doped Spiro-MeOTAD

film before (left) and after burn-in (right). The annealed film shows chain-like structures, leading to the irreversible

morphology degradation. (d) The morphology of a VNPB-MoO3 film before (left) and after burn-in (right). No visible

morphology evolution can be observed. (e) The evolution of perovskite content in the device active layer, tracked

using the PbI2 peak (•) and perovskite peak (*) in XRD measurements. Devices are tested after storage in air (70%

RH, dark) for 10 days and 30 days. PbI2 peak of the perovskite film in a Spiro-MeOTAD device (left) emerges after

10 days (grey) and dominates the perovskite peak after 30 days (blue), indicating severe decomposition of the

perovskite phase. In contrast, the perovskite layer is well protected by the VNPB-MoO3 film (right) and shows

negligible PbI2 signal even after 30 days (red).

I used X-ray diffraction (XRD) to explore the evolution of the perovskite active layer in devices

placed under stress via the introduction of moisture (70% RH) combined with elevated

0.0 0.2 0.4 0.6 0.8 1.0 1.2

-10

0

10

20

Control

Before

After

Cu

rre

nt d

en

sity (

mA

cm

-2)

Voltage (V)

BeforeAnnealing (110 ˚C, 1 hour)

Co

ntr

ol

Cro

sslin

ked

a b

c

d

0.0 0.2 0.4 0.6 0.8 1.0 1.2

-10

0

10

20

Crosslinked

Before

After

Cu

rre

nt d

en

sity (

mA

cm

-2)

Voltage (V)

10 15 20

Crosslinked

Perovskite

10 15 20

XR

D c

ou

nts

(a

.u.)

2deg

Control

PbI2

**

After 30 days

After 10 days

PerovskiteeAfter

Page 64: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

52

temperatures. The degree of perovskite degradation was quantified by the ratio of PbI2 peaks to

perovskite peaks. In control devices with spiro-MeOTAD as the top contact, degradation is

noticeable within 10 days and becomes significant after 30 days (Figure 5-6e, left). After 30 days

the PbI2 peak exceeds the perovskite signal, and the film is visibly yellow.

In contrast, following 30 days of moist heat, the perovskite in devices covered by the crosslinked

layer did not change within measurement uncertainty (Figure 5-6e, right). This finding further

confirms that the crosslinked hole transport layer, coupled with dense inorganic metal oxide in a

double-layer fashion, provides the perovskite with superior physical protection.

I also investigated the potential to adapt the new device structure in the direction of enabling

multi-junction cells. I measured the impact of subsequent solvent exposure (Figure 5-7),

applying a polar solvent (methanol) often used in follow-on layer fabrication. The conventional

spiro-MeOTAD device was much degraded as seen in its bandedge absorption change (Figure

5-8a, left). When the same device is exposed to chlorobenzene, the sandwiched inorganic spiro-

MeOTAD layer dissolves, producing irreversible loss of device structure and morphology

(Figure 5-8b, upper). I concluded that the conventional perovskite materials stack is vulnerable

even when subjected to nominally orthogonal solvents.

In contrast, the device covered with a crosslinked HTL retains the active material and the device

structure when exposed to both polar and nonpolar solvents (Figure 5-8a, right; Figure 5-8b,

downside). The enhanced resistance to heat, moisture, and follow-on solvent-based processing

not only benefit the single cell, but also open avenues for fabricating multi-junction devices atop

the perovskite.

Page 65: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

53

Figure 5-7. Assessment of device evolution under the external solvent attack. For the sake of fair comparison and

simulation of multijunction fabrication scenario, the same MoO3 layer and ZnO layer were deposited on top of (a)

Spiro-covered device and (b) crosslinked VNPB covered device, using high vacuum deposition methods. Two devices

(a) and (b), are soaked in both polar solvent (methanol) and nonpolar solvent (chlorobenzene) for 30s to investigate

the material or morphology evolution.

Figure 5-8. Evolution of material and morphology under the external solvent attack. (a) Evolution of perovskite

active layer after devices soaked in polar solvent (Methanol) for 30s. [The device layer configuration and test method

are shown in Figure 5-7.] The device using a Spiro hole transport layer (a, left) shows serious degradation of the

perovskite absorption edge, before (grey) and after (blue) methanol soaking. The device using a crosslinked hole

transport material (a, right) shows identical perovskite absorption before (grey) and after (red) solvent soaking. (b)

Evolution of top surface morphology after devices soaked in nonpolar solvent (chlorobenzene) for 30s. In device using

FTO on Glass

TiO2/PCBM

Perovskite

Crosslinked VNPBMoO3 (thermal evap)

ZnO (Sputter deposit, 50nm)

FTO on Glass

TiO2/PCBM

Perovskite

Doped Spiro-OMeTAD

MoO3 (thermal evap)

ZnO (Sputter deposit, 50nm)

Solvent soaking:Methanol; Chlorobenzene

Solvent soaking:Methanol; Chlorobenzene

a b

Control Crosslinked HETC

500 600 700 800

Spiro

Before

After

Ab

so

rptio

n (

a.u

.)

wavelength (nm)500 600 700 800

Crosslinked

Before

After

Methanol soaking

100µm

Chlorobenzene soaking Before After

Co

ntr

ol

Cro

sslin

ked

a b

Page 66: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

54

Spiro (b, upper), the top morphology is flat (b, upper left) before solvent soaking and becomes cracked across the

whole area of device after solvent soaking (b, upper right), due to the dissolution and reconstruction of the underlying

Spiro layer. In the device using the crosslinking VNPB material (b, down), the top surface morphology is kept

essentially the same before (b, down left) and after (b, down right) solvent soaking, due to the underlying insoluble

hole transport material.

5.7 Conclusions

In summary, I demonstrated a new methodology for hole extraction on top of planar perovskite

solar cells. The crosslinked organic hole transport material, coupled with an inorganic metal

oxide, provides an insoluble and inert physical protection layer, combined with high conductance

for hole extraction. This enabled device performance that is stable for longer durations and under

more intense external stresses than in many prior reports.

Page 67: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

55

Chapter 6

Conclusions

6.1 Summary and Impact

The research in this thesis was geared to providing new material and device designs to achieve

high efficiency perovskite solar cells. The focus was on resolving the problems – hysteresis and

instability – specific to this emerging solar PV system. By performing first-principles

simulations (DFT) and optoelectronic simulations, I was able to identify engineering routes at

both the materials and device levels. By gaining control over the material chemistry throughout

the interfaces within the photon-absorber and charge-extraction contact, I was then able to create

novel architectures that contributed to improved performance.

The work began with clarifying fundamental aspects underlying the impact of growth conditions

on the performance of perovskite films (Chapter 3). It revealed delocalization of the electronic

states within the local nanocrystal surfaces that preserves the integrity of the bulk bandgap. The

DFT-based analysis of defect formation energies identified the key defects (Pb atom substituted

by I) and indicated that films grown under iodine-rich conditions are prone to a high density of

deep electronic traps (recombination centers). This finding motivated the exploration of a new

precursor (lead acetate) for device-quality films. Insight into defect physics have spawned a

broad range of new perovskite growth processing reported in the wider literature.

The work achieved success in reducing current hysteresis and instability in planar cells through

the materials engineering of electron-extraction interfaces and grain boundaries (Chapter 4). I

reported the first perovskite-PCBM hybrid solid and found that the PCBM throughout the grain

boundaries and electron-extraction interfaces suppresses the hysteresis that has plagued planar

perovskite devices. I then conducted in-depth material characterizations and DFT simulations

that revealed the PCBM-perovskite interaction: the PCBM passivates the key PbI3- antisite

defects during perovskite self-assembly.

Using conductive AFM studies, I revealed the memristive properties of perovskite films and

identified the major cause of hysteresis to be ionic migration under electric fields in films, an

Page 68: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

56

initial step along this research front. This proposed mechanism of hysteresis now resonates with

experimental and theoretical results obtained in subsequent research reported in the literature.

I closed with the engineering of the exposed hole-extraction interface on the perovskite top

surface, which is of particular importance for further improving device stability and performance

simultaneously (Chapter 5). I developed the first crosslinked hole-extraction top contact to

obviate in situ degradation of the underlying perovskite. The new crosslinked hole-transport

medium produces an insoluble and heat-resistant materials stack atop the perovskite that is band-

aligned with the perovskite. I also found that hole-extracting contacts, which rely on chemical

doping, were the weak link from a stability perspective in the best-performing perovskite cells. I

therefore employed a new remote-doping strategy to induce the needed work function and free

carrier density. The resultant family of devices is hysteresis-free, with fill factors exceeding 80%

and with excellent resilience to thermal stresses that exceed 100˚C, conditions under which

conventionally-contacted devices fail. The devices are also resistant to stresses produced by

moisture and solvents that cause conventional devices to decompose.

My methodology represents a new strategy for constructing transparent, highly-conductive,

thermally-stable and solvent-resistant top contacts in perovskite devices. This top contact is the

base to build a second cell in a multi-junction configuration, paving the way for broader light

absorption and higher efficiency potential.

The work in this thesis directly resulted in the publication of 3 peer-reviewed journal articles that

have been cited >300 times during the past 2 years.

6.2 Outlook for perovskite solids and PV

6.2.1 Maximum efficiency

During the course of the work reported in this thesis, advances in perovskite PV continued, with

certified efficiencies (not stabilized) above 20% achieved after only 5 years of worldwide

research efforts126,151,152. There, however, is still room for further efficiency improvement toward

the Shockley–Queisser limit (VOC=1.32V; JSC=25mA cm-2; FF=90.5%; PCE=30% for a

semiconductor with bandgap Eg=1.6 eV). The relative current fraction is ~0.88 (Figure 6-1). The

current loss comes primarily from front reflection and parasitic absorption in the hole-extraction

layer and back contact. Compared with c-Si and GaAs solar cells, the voltage and fill factor loss

Page 69: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

57

in the presently-best perovskite PV is particularly high, because of the charge carrier

recombination paths in the absorber and at the interfaces, the shunting paths in non-ideal films

and carrier-extraction contacts, and resistive losses due to non-ideal contacts. From this analysis,

there is still a significant opportunity for further progress on material processing and interface

engineering for better light absorption and carrier management26.

Another research front lies in developing smaller bandgap perovskites which are desirable for

achieving a higher Shockley-Queisser efficiency limit. Highly efficient larger bandgap

perovskites are also of interest, valuable as the front cell in a multi-junction cell, such as a

perovskite-Si tandem cell with a potential efficiency that exceeds 30%.

Figure 6-1. Fraction of Shockley-Queisser detailed-balance limit for voltage and current achieved by record

cells. The lines crossing some data points indicate the uncertainty of bandgap of the record cell. ηSQ indicates the

maximum efficiency following the S-Q model. Figure reproduced from ref. 26. Copyright 2016 by the American

Association for the Advancement of Science.

6.2.2 Long-term stability

Despite the exciting efficiencies tested in labs, today’s best perovskite solar cells are known to

degrade within days under standard operating conditions. This is the greatest gap to commercial

cells that are required to work for more than 20 years. The origins of instabilities and hysteresis

are not totally understood yet, and should be kept as a topic of active research, even though UV

photo-reduction, water reaction, and ionic migration from defects have already been documented

perovskite

mc-Si

CdTe

CIGS

InP

GaInP

GaAs

c-Si1.00

0.95

0.90

0.85

0.80

0.75

0.70

0.65

0.6010.90.80.70.5 0.6

Carrier management

Ligh

t m

anag

em

en

t

v x f (FF VOC / FFSQ VSQ)

j(J SC

/ J SQ

)

Page 70: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

58

as mechanisms. Developing electron- and hole-extraction materials and contact designs for long-

term operation in perovskite devices is crucial. Finding new perovskite materials with stability

advantages without efficiency compromises will represent a big breakthrough in the future.

Compared with CdTe and GaAs solar cells, toxicity associated with Pb is more serious because

of the higher water solubility of perovskite. Therefore, intense research on perovskite materials

free of toxic elements is also of great worth.

Making big steps forward in perovskite photovoltaics relies on a coordinated international and

interdisciplinary effort. The urgent need for high efficiencies and low costs in photovoltaics

creates a powerful motivation to continue pursuing rapid advances. The further success of

solution-processed perovskite thin film photovoltaics will enable even greater penetration of

renewable electricity into our energy system and daily life in the future, spanning solar farms,

building-integrated systems, and mobile electronics.

Page 71: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

59

References

1. Jeff Taso, Nate Lewis & George Crabtree. Solar FAQs. (U.S. Department of Energy, 2006).

2. Renewables 2016 Global Status Report. REN21 (2016). Available at: http://www.ren21.net/status-of-

renewables/global-status-report/.

3. Nathan S. Lewis, George Crabtree, Arthur J. Nozik, Michael R. Wasielewski & Paul Alivisatos. Basic Research

Needs in Solar Energy Utilization. (U.S. Department of Energy, 2005).

4. Green, M. A., Emery, K., Hishikawa, Y., Warta, W. & Dunlop, E. D. Solar cell efficiency tables (Version 45):

Solar cell efficiency tables. Prog. Photovolt. Res. Appl. 23, 1–9 (2015).

5. Panasonic HIT(R) Solar Cell Achieves World’s Highest Energy Conversion Efficiency of 25.6% at Research

Level | Headquarters News | Panasonic Newsroom Global. (2014). Available at:

http://news.panasonic.com/global/press/data/2014/04/en140410-4/en140410-4.html. (Accessed: 3rd November

2016)

6. Kayes, B. M. et al. 27.6% Conversion efficiency, a new record for single-junction solar cells under 1 sun

illumination. in 2011 37th IEEE Photovoltaic Specialists Conference (PVSC) 000004–000008 (2011).

doi:10.1109/PVSC.2011.6185831

7. First Solar Achieves Efficiency, Durability Milestones (NASDAQ:FSLR). Available at:

http://investor.firstsolar.com/releasedetail.cfm?ReleaseID=895118. (Accessed: 3rd November 2016)

8. Jackson, P. et al. Properties of Cu(In,Ga)Se 2 solar cells with new record efficiencies up to 21.7%: Properties of

Cu(In,Ga)Se 2 solar cells with new record efficiencies up to 21.7%. Phys. Status Solidi RRL - Rapid Res. Lett. 9,

28–31 (2015).

9. O’Regan, B. & Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films.

Nature 353, 737–740 (1991).

10. Mathew, S. et al. Dye-sensitized solar cells with 13% efficiency achieved through the molecular engineering of

porphyrin sensitizers. Nat. Chem. 6, 242–247 (2014).

11. R. Komiya et al. Improvement of the conversion efficiency of a monolithic type dye-sensitized solar cell

module. in (2011).

12. Shaheen, S. E. et al. 2.5% efficient organic plastic solar cells. Appl. Phys. Lett. 78, 841–843 (2001).

13. Yu, G., Gao, J., Hummelen, J. C., Wudl, F. & Heeger, A. J. Polymer Photovoltaic Cells: Enhanced Efficiencies

via a Network of Internal Donor-Acceptor Heterojunctions. Science 270, 1789–1791 (1995).

14. Koleilat, G. I. et al. Efficient, Stable Infrared Photovoltaics Based on Solution-Cast Colloidal Quantum Dots.

ACS Nano 2, 833–840 (2008).

15. Lan, X. et al. 10.6% Certified Colloidal Quantum Dot Solar Cells via Solvent-Polarity-Engineered Halide

Passivation. Nano Lett. 16, 4630–4634 (2016).

16. Sargent, E. H. Colloidal quantum dot solar cells. Nat. Photonics 6, 133–135 (2012).

17. Todorov, T. & Mitzi, D. B. Direct Liquid Coating of Chalcopyrite Light-Absorbing Layers for Photovoltaic

Devices. Eur. J. Inorg. Chem. 2010, 17–28 (2010).

18. Wang, W. et al. Device Characteristics of CZTSSe Thin-Film Solar Cells with 12.6% Efficiency. Adv. Energy

Mater. 4, 1301465 (2014).

19. Haight, R. et al. Band alignment at the Cu2ZnSn(SxSe1−x)4/CdS interface. Appl. Phys. Lett. 98, 253502

(2011).

20. Solar Spectral Irradiance: Air Mass 1.5. Available at: http://rredc.nrel.gov/solar/spectra/am1.5/. (Accessed: 3rd

November 2016)

21. Shockley, W. & Queisser, H. J. Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. J. Appl. Phys.

32, 510 (1961).

22. M. A. Green. Third Generation Photovoltaics. 12, (Springer Berlin Heidelberg, 2006).

23. King, R. R. Multijunction cells: Record breakers. Nat. Photonics 2, 284–286 (2008).

24. Sharp Develops Concentrator Solar Cell with World’s Highest Conversion Efficiency of 44.4% | Press Releases

| Sharp Global. Available at: http://www.sharp-world.com/corporate/news/130614.html. (Accessed: 3rd

November 2016)

25. Graetzel, M., Janssen, R. A. J., Mitzi, D. B. & Sargent, E. H. Materials interface engineering for solution-

processed photovoltaics. Nature 488, 304–312 (2012).

26. Polman, A., Knight, M., Garnett, E. C., Ehrler, B. & Sinke, W. C. Photovoltaic materials: Present efficiencies

and future challenges. Science 352, aad4424 (2016).

Page 72: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

60

27. Nazeeruddin, M. K., Baranoff, E. & Grätzel, M. Dye-sensitized solar cells: A brief overview. Sol. Energy 85,

1172–1178 (2011).

28. Yella, A. et al. Porphyrin-Sensitized Solar Cells with Cobalt (II/III)–Based Redox Electrolyte Exceed 12

Percent Efficiency. Science 334, 629–634 (2011).

29. Chen, H.-Y. et al. Polymer solar cells with enhanced open-circuit voltage and efficiency. Nat. Photonics 3,

649–653 (2009).

30. O’Malley, K. M., Li, C.-Z., Yip, H.-L. & Jen, A. K.-Y. Enhanced Open-Circuit Voltage in High Performance

Polymer/Fullerene Bulk-Heterojunction Solar Cells by Cathode Modification with a C60 Surfactant. Adv.

Energy Mater. 2, 82–86 (2012).

31. Ip, A. H. et al. Hybrid passivated colloidal quantum dot solids. Nat. Nanotechnol. 7, 577–582 (2012).

32. Chen, S., Gong, X. G., Walsh, A. & Wei, S.-H. Defect physics of the kesterite thin-film solar cell absorber

Cu2ZnSnS4. Appl. Phys. Lett. 96, 21902 (2010).

33. Cao, Q. et al. Defects in Cu(In,Ga)Se2 Chalcopyrite Semiconductors: A Comparative Study of Material

Properties, Defect States, and Photovoltaic Performance. Adv. Energy Mater. 1, 845–853 (2011).

34. Kojima, A., Teshima, K., Shirai, Y. & Miyasaka, T. Organometal Halide Perovskites as Visible-Light

Sensitizers for Photovoltaic Cells. J. Am. Chem. Soc. 131, 6050–6051 (2009).

35. Jeon, N. J. et al. Solvent engineering for high-performance inorganic–organic hybrid perovskite solar cells. Nat.

Mater. 13, 897–903 (2014).

36. Goldschmidt, V. M. Die Gesetze der Krystallochemie. Naturwissenschaften 14, 477–485 (1926).

37. Peña, M. A. & Fierro, J. L. G. Chemical Structures and Performance of Perovskite Oxides. Chem. Rev. 101,

1981–2018 (2001).

38. Stoumpos, C. C., Malliakas, C. D. & Kanatzidis, M. G. Semiconducting Tin and Lead Iodide Perovskites with

Organic Cations: Phase Transitions, High Mobilities, and Near-Infrared Photoluminescent Properties. Inorg.

Chem. 52, 9019–9038 (2013).

39. Cheng, Z. & Lin, J. Layered organic–inorganic hybrid perovskites: structure, optical properties, film

preparation, patterning and templating engineering. CrystEngComm 12, 2646–2662 (2010).

40. Mitzi, D. B., Wang, S., Feild, C. A., Chess, C. A. & Guloy, A. M. Conducting Layered Organic-inorganic

Halides Containing <110>-Oriented Perovskite Sheets. Science 267, 1473–1476 (1995).

41. McGehee, M. D. Perovskite solar cells: Continuing to soar. Nat. Mater. 13, 845–846 (2014).

42. Grätzel, M. The light and shade of perovskite solar cells. Nat. Mater. 13, 838–842 (2014).

43. Green, M. A., Ho-Baillie, A. & Snaith, H. J. The emergence of perovskite solar cells. Nat. Photonics 8, 506–

514 (2014).

44. De Wolf, S. et al. Organometallic Halide Perovskites: Sharp Optical Absorption Edge and Its Relation to

Photovoltaic Performance. J. Phys. Chem. Lett. 5, 1035–1039 (2014).

45. Dong, Q. et al. Electron-hole diffusion lengths >175 μm in solution grown CH3NH3PbI3 single crystals.

Science aaa5760 (2015). doi:10.1126/science.aaa5760

46. Shi, D. et al. Low trap-state density and long carrier diffusion in organolead trihalide perovskite single crystals.

Science 347, 519–522 (2015).

47. Xing, G. et al. Long-Range Balanced Electron- and Hole-Transport Lengths in Organic-Inorganic

CH3NH3PbI3. Science 342, 344–347 (2013).

48. Buin, A. et al. Materials Processing Routes to Trap-Free Halide Perovskites. Nano Lett. (2014).

doi:10.1021/nl502612m

49. Stranks, S. D. et al. Electron-Hole Diffusion Lengths Exceeding 1 Micrometer in an Organometal Trihalide

Perovskite Absorber. Science 342, 341–344 (2013).

50. D’Innocenzo, V. et al. Excitons versus free charges in organo-lead tri-halide perovskites. Nat. Commun. 5,

(2014).

51. Lin, Q., Armin, A., Nagiri, R. C. R., Burn, P. L. & Meredith, P. Electro-optics of perovskite solar cells. Nat.

Photonics 9, 106–112 (2015).

52. Kim, H.-S. et al. Lead Iodide Perovskite Sensitized All-Solid-State Submicron Thin Film Mesoscopic Solar

Cell with Efficiency Exceeding 9%. Sci. Rep. 2, (2012).

53. Lee, M. M., Teuscher, J., Miyasaka, T., Murakami, T. N. & Snaith, H. J. Efficient Hybrid Solar Cells Based on

Meso-Superstructured Organometal Halide Perovskites. Science 338, 643–647 (2012).

54. Burschka, J. et al. Sequential deposition as a route to high-performance perovskite-sensitized solar cells. Nature

499, 316–319 (2013).

55. Docampo, P., Ball, J. M., Darwich, M., Eperon, G. E. & Snaith, H. J. Efficient organometal trihalide perovskite

planar-heterojunction solar cells on flexible polymer substrates. Nat. Commun. 4, (2013).

Page 73: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

61

56. Malinkiewicz, O. et al. Perovskite solar cells employing organic charge-transport layers. Nat. Photonics 8, 128–

132 (2014).

57. Deschler, F. et al. High Photoluminescence Efficiency and Optically Pumped Lasing in Solution-Processed

Mixed Halide Perovskite Semiconductors. J. Phys. Chem. Lett. 5, 1421–1426 (2014).

58. Tan, Z.-K. et al. Bright light-emitting diodes based on organometal halide perovskite. Nat. Nanotechnol. 9,

687–692 (2014).

59. Habisreutinger, S. N. et al. Carbon Nanotube/Polymer Composites as a Highly Stable Hole Collection Layer in

Perovskite Solar Cells. Nano Lett. 14, 5561–5568 (2014).

60. Leijtens, T. et al. Overcoming ultraviolet light instability of sensitized TiO2 with meso-superstructured

organometal tri-halide perovskite solar cells. Nat. Commun. 4, (2013).

61. Ito, S., Tanaka, S., Manabe, K. & Nishino, H. Effects of Surface Blocking Layer of Sb2S3 on Nanocrystalline

TiO2 for CH3NH3PbI3 Perovskite Solar Cells. J. Phys. Chem. C 118, 16995–17000 (2014).

62. Li, W. et al. Montmorillonite as bifunctional buffer layer material for hybrid perovskite solar cells with

protection from corrosion and retarding recombination. J. Mater. Chem. A 2, 13587–13592 (2014).

63. Abate, A. et al. Lithium salts as ‘redox active’ p-type dopants for organic semiconductors and their impact in

solid-state dye-sensitized solar cells. Phys. Chem. Chem. Phys. 15, 2572–2579 (2013).

64. Bailie, C. D., Unger, E. L., Zakeeruddin, S. M., Grätzel, M. & McGehee, M. D. Melt-infiltration of spiro-

OMeTAD and thermal instability of solid-state dye-sensitized solar cells. Phys. Chem. Chem. Phys. PCCP 16,

4864–4870 (2014).

65. Snaith, H. J. et al. Anomalous Hysteresis in Perovskite Solar Cells. J. Phys. Chem. Lett. 5, 1511–1515 (2014).

66. Kim, H.-S. & Park, N.-G. Parameters Affecting I–V Hysteresis of CH3NH3PbI3 Perovskite Solar Cells: Effects

of Perovskite Crystal Size and Mesoporous TiO2 Layer. J. Phys. Chem. Lett. 5, 2927–2934 (2014).

67. Unger, E. L. et al. Hysteresis and transient behavior in current–voltage measurements of hybrid-perovskite

absorber solar cells. Energy Environ. Sci. 7, 3690–3698 (2014).

68. Abate, A. et al. Supramolecular Halogen Bond Passivation of Organic–Inorganic Halide Perovskite Solar Cells.

Nano Lett. 14, 3247–3254 (2014).

69. Agiorgousis, M. L., Sun, Y.-Y., Zeng, H. & Zhang, S. Strong Covalency-Induced Recombination Centers in

Perovskite Solar Cell Material CH 3 NH 3 PbI 3. J. Am. Chem. Soc. 136, 14570–14575 (2014).

70. Suarez, B. et al. Recombination Study of Combined Halides (Cl, Br, I) Perovskite Solar Cells. J. Phys. Chem.

Lett. 5, 1628–1635 (2014).

71. Tress, W. et al. Understanding the rate-dependent J-V hysteresis, slow time component, and aging in

CH3NH3PbI3 perovskite solar cells: the role of a compensated electric field. Energy Environ. Sci. 8, (2015).

72. Shao, Y., Xiao, Z., Bi, C., Yuan, Y. & Huang, J. Origin and elimination of photocurrent hysteresis by fullerene

passivation in CH3NH3PbI3 planar heterojunction solar cells. Nat. Commun. 5, (2014).

73. Quilettes, D. W. de et al. Impact of microstructure on local carrier lifetime in perovskite solar cells. Science

348, 683–686 (2015).

74. Mosconi, E., Amat, A., Nazeeruddin, M. K., Grätzel, M. & De Angelis, F. First-Principles Modeling of Mixed

Halide Organometal Perovskites for Photovoltaic Applications. J. Phys. Chem. C 117, 13902–13913 (2013).

75. Wang, Y. et al. Density functional theory analysis of structural and electronic properties of orthorhombic

perovskite CH 3 NH 3 PbI 3. Phys Chem Chem Phys 16, 1424–1429 (2014).

76. Lindblad, R. et al. Electronic Structure of TiO2/CH3NH3PbI3 Perovskite Solar Cell Interfaces. J. Phys. Chem.

Lett. 5, 648–653 (2014).

77. Haruyama, J., Sodeyama, K., Han, L. & Tateyama, Y. Termination Dependence of Tetragonal CH3NH3PbI3

Surfaces for Perovskite Solar Cells. J. Phys. Chem. Lett. 5, 2903–2909 (2014).

78. Liu, M., Johnston, M. B. & Snaith, H. J. Efficient planar heterojunction perovskite solar cells by vapour

deposition. Nature 501, 395–398 (2013).

79. Yin, W.-J., Shi, T. & Yan, Y. Unusual defect physics in CH3NH3PbI3 perovskite solar cell absorber. Appl.

Phys. Lett. 104, 63903 (2014).

80. Noh, J. H., Im, S. H., Heo, J. H., Mandal, T. N. & Seok, S. I. Chemical Management for Colorful, Efficient, and

Stable Inorganic–Organic Hybrid Nanostructured Solar Cells. Nano Lett. 13, 1764–1769 (2013).

81. Niu, G. et al. Study on the stability of CH3NH3PbI3 films and the effect of post-modification by aluminum

oxide in all-solid-state hybrid solar cells. J. Mater. Chem. A 2, 705–710 (2013).

82. Sun, S. et al. The origin of high efficiency in low-temperature solution-processable bilayer organometal halide

hybrid solar cells. Energy Environ. Sci. 7, 399–407 (2013).

83. Janotti, A. & Van de Walle, C. G. Native point defects in ZnO. Phys. Rev. B 76, 165202 (2007).

84. Walle, C. G. V. de & Neugebauer, J. First-principles calculations for defects and impurities: Applications to III-

nitrides. J. Appl. Phys. 95, 3851–3879 (2004).

Page 74: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

62

85. Stampfl, C. & Van de Walle, C. G. Theoretical investigation of native defects, impurities, and complexes in

aluminum nitride. Phys. Rev. B 65, 155212 (2002).

86. Farrell, D. E. & Wolverton, C. First-principles study of point defects under varied chemical potentials in

Li${}_{4}$BN${}_{3}$H${}_{10}$. Phys. Rev. B 85, 174102 (2012).

87. Boukhvalov, D. W. & Solovyev, I. V. Defects of the crystal structure and Jahn-Teller distortion in

${\text{BiMnO}}_{3}$. Phys. Rev. B 82, 245101 (2010).

88. Derenzo, S. E. et al. Experimental and theoretical studies of donor–acceptor scintillation from PbI2. J. Lumin.

134, 28–34 (2013).

89. Kim, J., Lee, S.-H., Lee, J. H. & Hong, K.-H. The Role of Intrinsic Defects in Methylammonium Lead Iodide

Perovskite. J. Phys. Chem. Lett. 5, 1312–1317 (2014).

90. Lany, S. & Zunger, A. Anion vacancies as a source of persistent photoconductivity in II-VI and chalcopyrite

semiconductors. Phys. Rev. B 72, 35215 (2005).

91. Shockley, W. & Read, W. T. Statistics of the Recombinations of Holes and Electrons. Phys. Rev. 87, 835–842

(1952).

92. Knop, Osvald et al. Alkylammonium lead halides. Part 2. CH3NH3PbX3 (X = chlorine, bromine, iodine)

perovskites: cuboctahedral halide cages with isotropic cation reorientation. Can. J. Chem. 68, 412–422 (1990).

93. Zhang, H., Jang, J., Liu, W. & Talapin, D. V. Colloidal Nanocrystals with Inorganic Halide, Pseudohalide, and

Halometallate Ligands. ACS Nano 8, 7359–7369 (2014).

94. Zhitomirsky, D., Voznyy, O., Hoogland, S. & Sargent, E. H. Measuring Charge Carrier Diffusion in Coupled

Colloidal Quantum Dot Solids. ACS Nano 7, 5282–5290 (2013).

95. Liu, D. & Kelly, T. L. Perovskite solar cells with a planar heterojunction structure prepared using room-

temperature solution processing techniques. Nat. Photonics 8, 133–138 (2014).

96. Zhang, W. et al. Ultrasmooth organic–inorganic perovskite thin-film formation and crystallization for efficient

planar heterojunction solar cells. Nat. Commun. 6, 6142 (2015).

97. Wehrenfennig, C., Eperon, G. E., Johnston, M. B., Snaith, H. J. & Herz, L. M. High Charge Carrier Mobilities

and Lifetimes in Organolead Trihalide Perovskites. Adv. Mater. 26, 1584–1589 (2014).

98. You, J. et al. Low-Temperature Solution-Processed Perovskite Solar Cells with High Efficiency and Flexibility.

ACS Nano 8, 1674–1680 (2014).

99. Gonzalez-Pedro, V. et al. General Working Principles of CH3NH3PbX3 Perovskite Solar Cells. Nano Lett. 14,

888–893 (2014).

100. Ball, J. M., Lee, M. M., Hey, A. & Snaith, H. J. Low-temperature processed meso-superstructured to thin-film

perovskite solar cells. Energy Environ. Sci. 6, 1739–1743 (2013).

101. Du, M. H. Efficient carrier transport in halide perovskites: theoretical perspectives. J. Mater. Chem. A 2, 9091–

9098 (2014).

102. Xiao, M. et al. A Fast Deposition-Crystallization Procedure for Highly Efficient Lead Iodide Perovskite Thin-

Film Solar Cells. Angew. Chem. Int. Ed. 53, 9898–9903 (2014).

103. Zhou, H. et al. Interface engineering of highly efficient perovskite solar cells. Science 345, 542–546 (2014).

104. Guldi, D. M. & Prato, M. Excited-State Properties of C60 Fullerene Derivatives. Acc. Chem. Res. 33, 695–703

(2000).

105. Li, C.-Z. et al. Doping of Fullerenes via Anion-Induced Electron Transfer and Its Implication for Surfactant

Facilitated High Performance Polymer Solar Cells. Adv. Mater. 25, 4425–4430 (2013).

106. Weber, C. D., Bradley, C. & Lonergan, M. C. Solution phase n-doping of C60 and PCBM using

tetrabutylammonium fluoride. J. Mater. Chem. A 2, 303–307 (2013).

107. Popescu, L. M., Hof, P. van ’t, Sieval, A. B., Jonkman, H. T. & Hummelen, J. C. Thienyl analog of 1-(3-

methoxycarbonyl)propyl-1-phenyl-[6,6]-methanofullerene for bulk heterojunction photovoltaic devices in

combination with polythiophenes. Appl. Phys. Lett. 89, 213507 (2006).

108. Kumar, P. P. & Yashonath, S. Ionic conduction in the solid state. J. Chem. Sci. 118, 135–154 (2006).

109. Lehovec, K. Space‐Charge Layer and Distribution of Lattice Defects at the Surface of Ionic Crystals. J. Chem.

Phys. 21, 1123–1128 (1953).

110. Chua, L. O. Memristor-The missing circuit element. IEEE Trans. Circuit Theory 18, 507–519 (1971).

111. Strukov, D. B., Snider, G. S., Stewart, D. R. & Williams, R. S. The missing memristor found. Nature 453, 80–

83 (2008).

112. Xiao, Z. et al. Giant switchable photovoltaic effect in organometal trihalide perovskite devices. Nat. Mater.

advance online publication, (2014).

113. Guo, X. et al. Polymer solar cells with enhanced fill factors. Nat. Photonics 7, 825–833 (2013).

114. Liu, P. et al. Interfacial electronic structure at the CH3NH3PbI3/MoOx interface. Appl. Phys. Lett. 106, 193903

(2015).

Page 75: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

63

115. Kaltenbrunner, M. et al. Flexible high power-per-weight perovskite solar cells with chromium oxide-metal

contacts for improved stability in air. Nat. Mater. 14, 1032–1039 (2015).

116. Kato, Y. et al. Silver Iodide Formation in Methyl Ammonium Lead Iodide Perovskite Solar Cells with Silver

Top Electrodes. Adv. Mater. Interfaces 2, n/a-n/a (2015).

117. Li, X. et al. Improved performance and stability of perovskite solar cells by crystal crosslinking with

alkylphosphonic acid ω-ammonium chlorides. Nat. Chem. 7, 703–711 (2015).

118. Leijtens, T. et al. Stability of Metal Halide Perovskite Solar Cells. Adv. Energy Mater. 5, n/a-n/a (2015).

119. Mei, A. et al. A hole-conductor–free, fully printable mesoscopic perovskite solar cell with high stability.

Science 345, 295–298 (2014).

120. Jung, M.-C., Raga, S. R., Ono, L. K. & Qi, Y. Substantial improvement of perovskite solar cells stability by

pinhole-free hole transport layer with doping engineering. Sci. Rep. 5, 9863 (2015).

121. Mailoa, J. P. et al. A 2-terminal perovskite/silicon multijunction solar cell enabled by a silicon tunnel junction.

Appl. Phys. Lett. 106, 121105 (2015).

122. Kranz, L. et al. High-Efficiency Polycrystalline Thin Film Tandem Solar Cells. J. Phys. Chem. Lett. 6, 2676–

2681 (2015).

123. Jeon, N. J. et al. Compositional engineering of perovskite materials for high-performance solar cells. Nature

517, 476–480 (2015).

124. Im, J.-H., Jang, I.-H., Pellet, N., Grätzel, M. & Park, N.-G. Growth of CH3NH3PbI3 cuboids with controlled

size for high-efficiency perovskite solar cells. Nat. Nanotechnol. 9, 927–932 (2014).

125. Heo, J. H. et al. Efficient inorganic-organic hybrid heterojunction solar cells containing perovskite compound

and polymeric hole conductors. Nat. Photonics 7, 486–491 (2013).

126. Yang, W. S. et al. High-performance photovoltaic perovskite layers fabricated through intramolecular

exchange. Science 348, 1234–1237 (2015).

127. Cappel, U. B., Daeneke, T. & Bach, U. Oxygen-Induced Doping of Spiro-MeOTAD in Solid-State Dye-

Sensitized Solar Cells and Its Impact on Device Performance. Nano Lett. 12, 4925–4931 (2012).

128. Nguyen, W. H., Bailie, C. D., Unger, E. L. & McGehee, M. D. Enhancing the Hole-Conductivity of Spiro-

OMeTAD without Oxygen or Lithium Salts by Using Spiro(TFSI)2 in Perovskite and Dye-Sensitized Solar

Cells. J. Am. Chem. Soc. 136, 10996–11001 (2014).

129. Hawash, Z., Ono, L. K., Raga, S. R., Lee, M. V. & Qi, Y. Air-Exposure Induced Dopant Redistribution and

Energy Level Shifts in Spin-Coated Spiro-MeOTAD Films. Chem. Mater. 27, 562–569 (2015).

130. Pattanasattayavong, P. et al. Electric field-induced hole transport in copper(I) thiocyanate (CuSCN) thin-films

processed from solution at room temperature. Chem. Commun. 49, 4154–4156 (2013).

131. Ito, S. et al. Carbon-Double-Bond-Free Printed Solar Cells from TiO2/CH3NH3PbI3/CuSCN/Au: Structural

Control and Photoaging Effects. ChemPhysChem 15, 1194–1200 (2014).

132. Müller, C. D. et al. Multi-colour organic light-emitting displays by solution processing. Nature 421, 829–833

(2003).

133. Müller, D. C. et al. Efficient Blue Organic Light-Emitting Diodes with Graded Hole-Transport Layers.

ChemPhysChem 1, 207–211 (2000).

134. Jhuo, H.-J. et al. Inverted perovskite solar cells with inserted cross-linked electron-blocking interlayers for

performance enhancement. J. Mater. Chem. A 3, (2015).

135. Deng, Y. et al. Scalable fabrication of efficient organolead trihalide perovskite solar cells with doctor-bladed

active layers. Energy Environ. Sci. 8, (2015).

136. Hayashi, N., Nishio, R. & TAKADA, S. Composition, film using the composition, charge transport layer,

organic electroluminescence device, and method for forming charge transport layer. (2012).

137. Chiang, T.-Y. et al. Functional p-Type, Polymerized Organic Electrode Interlayer in CH3NH3PbI3

Perovskite/Fullerene Planar Heterojunction Hybrid Solar Cells. ACS Appl. Mater. Interfaces 7, 24973–24981

(2015).

138. Lee, J.-H. & Kim, J.-J. Interfacial doping for efficient charge injection in organic semiconductors. Phys. Status

Solidi A 209, 1399–1413 (2012).

139. Kröger, M. et al. Role of the deep-lying electronic states of MoO3 in the enhancement of hole-injection in

organic thin films. Appl. Phys. Lett. 95, 123301 (2009).

140. Meyer, J. et al. Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications.

Adv. Mater. 24, 5408–5427 (2012).

141. Xu, J. et al. Perovskite-fullerene hybrid materials suppress hysteresis in planar diodes. Nat. Commun. 6, (2015).

142. Zhou, X. et al. Enhanced Hole Injection into Amorphous Hole-Transport Layers of Organic Light-Emitting

Diodes Using Controlled p-Type Doping. Adv. Funct. Mater. 11, 310–314 (2001).

Page 76: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

64

143. Leem, D.-S. et al. Low driving voltage and high stability organic light-emitting diodes with rhenium oxide-

doped hole transporting layer. Appl. Phys. Lett. 91, 11113 (2007).

144. Matsushima, T., Kinoshita, Y. & Murata, H. Formation of Ohmic hole injection by inserting an ultrathin layer

of molybdenum trioxide between indium tin oxide and organic hole-transporting layers. Appl. Phys. Lett. 91,

253504 (2007).

145. Zhao, Y. et al. Transition metal oxides on organic semiconductors. Org. Electron. 15, 871–877 (2014).

146. Ng, T.-W., Lo, M.-F., Yang, Q.-D., Fung, M.-K. & Lee, C.-S. Near-Infrared Electric Power Generation

Through Sub-Energy-Gap Absorption in an Organic–Inorganic Composite. Adv. Funct. Mater. 22, 3035–3042

(2012).

147. Yoo, S.-J. et al. Formation of perfect ohmic contact at indium tin oxide/N,N′-di(naphthalene-1-yl)-N,N′-

diphenyl-benzidine interface using ReO3. Sci. Rep. 4, (2014).

148. Matsushima, T. et al. Interfacial charge transfer and charge generation in organic electronic devices. Org.

Electron. 12, 520–528 (2011).

149. Burgelman, M., Nollet, P. & Degrave, S. Modelling polycrystalline semiconductor solar cells. Thin Solid Films

361–362, 527–532 (2000).

150. Wang, X. et al. Tandem colloidal quantum dot solar cells employing a graded recombination layer. Nat.

Photonics 5, 480–484 (2011).

151. Li, X. et al. A vacuum flash–assisted solution process for high-efficiency large-area perovskite solar cells.

Science aaf8060 (2016). doi:10.1126/science.aaf8060

152. NREL. Research Cell Record Efficiency Chart. Available at:

http://www.nrel.gov/pv/assets/images/efficiency_chart.jpg. (Accessed: 4th November 2016)

Page 77: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

65

Appendices

1 Methods

1.1 Anhydrate lead acetate protocol. 10 g lead (II) acetate trihydrate (Sigma-Aldrich,

99.99%) was dissolved in 10 ml of acetic acid anhydride at ~80˚C and distilled off at ~40

˚C under vacuum. The solid residue is anhydrate lead acetate.

1.2 Perovskite solution preparation. The dehydrated lead acetate (Pb(Ac)2) and

methylammonium iodide (MAI) (Dyesol, 99%+) were dissolved in DMF (N,N-

Dimethylformamide, Sigma-Aldrich, 99.9%) with the molar ratio 1:3 to form the

perovskite precursor solution. To obtain ultrathin films and thick films, we tune the

perovskite concentration between 0.2 M and 1 M.

1.3 Perovskite-PCBM hybrid solution preparation. PCBM (Nano-C, 99.5%) is mixed into

the perovskite solution. In typical procedure, the PCBM-perovskite weight ratios are

between 1:100 and 1:10. Specifically, PCBM can be dissolved into chlorobenzene first,

and then mixed with perovskite solution before spin coating. The solution is kept at 70˚C

before spinning. For low mixture ratio, the miscibility of mixture solution is good and can

be stabilized at room temperature; for high ratio mixtures approaching 1:10 and beyond,

the solution needs to be used quickly after mixing.

1.4 Planar perovskite film and device fabrication (Chapter 3 and 4). A thin TiO2 compact

layer was first formed on FTO substrate using magnetron sputtering (~50 nm, Kurt J.

Lesker, 99.9%) followed by a low-concentration TiCl4 treatment for interfacial contact

improvement: soak in 120 mM TiCl4 aqueous solution at 70˚C for 0.5 hour followed by

annealing at 500˚C for 0.5 hour. Perovskite-PCBM hybrid solid films were deposited on

pre-heated TiO2 substrate using spin-coating at 3000~5000 rpm for 60 seconds in a

nitrogen glovebox. During the spin-coating, the film turned to dark brown, implying that

the perovskite crystallization was almost done. The hybrid solid film was then heated for

10 minutes at 70˚C to remove the residual solvent. For a control planar heterojunction

device, pure perovskite solution is deposited on TiO2 substrate in the same way. No excess

acetate (Ac-) and methylammonium (MA-) were found in the final films. For bilayer

control devices, PCBM in chlorobenzene (~20 mg ml-1) was spin cast on a TiCl4-treated

TiO2 substrate and then annealed at 70˚C for 10 minutes before spin-coating the perovskite

on top. A thin PCBM layer (< 30 nm) between TiO2 and perovskite is formed. Hole

Page 78: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

66

transfer layer was deposited by spin-coating of Spiro-OMeTAD (Borun Chemical, 99%+)

solution following the doping procedure reported below in Appendix 1.6. Top contact was

50 nm thermally evaporated gold through the shadow mask under 10-7 torr vacuum using

an Angstrom Engineering deposition system.

1.5 Planar perovskite device fabrication (Chapter 5). A thin TiO2 compact layer is first

formed on FTO substrates using atomic layer deposition (ALD) (~10 nm, Cambridge

Nanotech Savannah S100) using tetrakis-dimethyl-amido titanium and H2O as precursors.

A low-concentration TiCl4 treatment is used for interfacial improvement. The substrates

are soaked in TiCl4 aqueous solution (120 mM, 70˚C) for 0.5 hour and then annealed at

500˚C for 0.5 hour. PCBM ([6,6]-phenyl-C61-butyric acid methyl ester, Nano-C, 99.5%)

in chlorobenzene (~20 mg ml-1) is spin cast on the TiO2 substrates and then annealed at

70˚C for 10 minutes before spin-coating the perovskite on top. Anhydrate lead acetate

(Pb(C2H3O2)2). and methylammonium iodide (MAI) (Dyesol, 99%+) are dissolved in DMF

(N,N-Dimethylformamide, Sigma-Aldrich, 99.9%) with the molar ratio 1:3 to form the

perovskite precursor solution (1~1.3 M) and kept at 70˚C. Perovskite precursor is mixed

with 20 ul PCBM in chlorobenzene (30 mg ml-1) and deposited by spin-casting

(3000~5000 rpm for 60 seconds) on pre-heated TiO2-PCBM substrates in a nitrogen

glovebox. The film is annealed at 75˚C for 5 minutes and then 100˚C for 15 minutes. For

control devices using chemical doping, the hole transfer layer was deposited by spin-

coating the mixture solution of Spiro-MeOTAD (Borun Chemical, 99%+), dopant Li-TFSI

(Bis(trifluoromethane)sulfonimide lithium salt) and additive tBP (4-tert-butylpyridine)

following below Appendix 1.6. For the interface doping devices, the hole transport layer

VNPB (Lumtec, 95%+) in anhydrous toluene (3 mg ml-1) is spin cast (3000~4000 rpm,

30s) on perovskite film, followed by processing for the thermal crosslinking (120˚C for

20minutes and 150˚C for 10 minutes). The interface doping layer is 10 nm MoO3,

evaporated under 10-7 torr vacuum (Angstrom Engineering deposition system). After that,

the samples are kept at 40 ˚C in the evaporation chamber for 10 minutes. The top metal

contact is gold (50 nm) deposited through a shadow mask. Encapsulation is done using the

UV cured epoxy (Ossila) in conjunction with a glass coverslip.

1.6 Spiro-MeOTAD doping protocol. Spiro-MeOTAD was dissolved in chlorobenzene (63

mg ml-1). Then tBP (4-tert-butylpyridine) was added as additive ( 20 μl ml-1). Dopant Li-

Page 79: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

67

TFSI (Bis(trifluoromethane)sulfonimide lithium salt) (170 mg ml-1 in acetonitrile) was

finally added into the prepared Spiro-MeOTAD solution (70 μl ml-1).

2 Characterizations

2.1 Steady-state photovoltaic performance and hysteresis-effect characterization. The

active area of devices is determined by an optical aperture (area 0.049 cm2) placed before

the device. The AM1.5 solar simulator (ScienceTech) is class A (<25% spectrum

mismatch) and the spectral mismatching factor was characterized using a Newport

calibrated reference Si solar cell. The spectral mismatching factor was used in every

reported performance. The illumination intensity on devices was calibrated using a Melles–

Griot power meter to be 1sun (100 mW cm− 2). The final accuracy of the solar-to-

electricity measurements was estimated to be ± 5%. Steady-state performance was

measured using a Keithley 2400 SourceMeter. A standard testing process is as follows:

first the steady-state open-circuit voltage VOC(t) is measured by fixing the current to zero;

then short-circuit current JSC(t) is measured by setting the voltage to zero; thirdly, the

forward- and reverse-scanning instantaneous J-V curves are measured with a scanning rate

of 0.2 V s-1 and the voltage of maximum power point (MPP) is determined. The J-V

voltage scanning range is 1.1~1.2 times the steady-state open-circuit voltage. The

hysteresis factor of J-V curves is quantified using Equation (5-1). Finally, the steady-state

power conversion efficiency (PCE(t)) is measured by setting the bias at the maximum

power point and tracking the output steady-state current for a certain duration. To avoid the

overestimation due to the hysteresis effect, the figure of merit of photovoltaic performance

is only determined by the steady-state efficiency.

2.2 External Quantum Efficiency (EQE) spectra is measured by aligning the cell to

monochromatic illumination (a 400W Xe lamp passing through a monochromator and

appropriate cut-off filters). The active area was defined by the optical aperture before the

cell, and the power was calibrated with UV-IR photodetectors (Newport 818-UV and

Newport 838-IR). A solar simulator at 1 sun intensity provided the light bias. The

monochromatic beam was chopped at 220Hz. The response of the cell was measured with

a pre-amplifier (Lakeshore) connected to a lock-in amplifier (Stanford Research 830) at

short circuit conditions.

2.3 Conductive atomic force microscope characterization (CAFM, Chapter 4). Scanning

probe microscopy experiments were carried out in a commercial ultrahigh-vacuum atomic

Page 80: Materials interface engineering in perovskite photovoltaics · Materials interface engineering in perovskite photovoltaics Jixian XU Doctor of Philosophy The Edward S. Rogers Sr

68

force microscope (UHV bean-deflection AFM, Omicron) using Cr/Pt-coated silicon

cantilevers (Budget Sensor, Multi75E-G). All the measurements were performed at a

background pressure of < 2 x 10–10 Torr after transferring the samples from ambient

without any additional treatment. Contact-mode AFM images and 2D current maps are

simultaneously obtained with the tip in contact with the surface (loading force ~1 nN)

applying fixed bias voltages. The I-V curves were acquired in the conductive AFM regime

from various locations of the sample surfaces applying a linear bias ramp with a rate of

~0.5 V s-1.

2.4 X-ray diffraction (XRD) measurements were performed at room temperature with a

Rigaku Miniflex 2-circle diffractometer operating in Bragg–Brentano scanning mode, with

angular resolution of 0.01 degrees and Cu-K radiation (0.154056 nm wavelength).

2.5 Ultraviolet photoelectron spectroscopy (UPS) was carried out using He I (21.22 eV)

photon lines from a discharge lamp.

2.6 XPS study on stoichiometry. X-ray photoelectron spectroscopy (XPS) is carried out using

a Thermo Scientific K-Alpha spectrometer. Core level spectra of Pb-4f, I-3d, O-1s, N-1s

and C-1s with a pass energy of 75 eV. The elemental composition was calculated based on

integrated counts of respective peaks. The curves were fitted using Gaussian functions with

1.5 eV FWHM. For comparison of different samples, all spectra were normalized to Pb

signal.

2.7 UV-Vis-IR absorption was measured using a PerkinElmer LAMBDA 950

Spectrophotometer.

2.8 Transient photoluminescence (PL) was carried out using TCSPC function of a HORIBA

FLuorolog-3 Spectrofluorometer. Samples were tested in N2 ambient.