transient infrared spectroscopy of charge …

145
The Pennsylvania State University The Graduate School Department of Chemistry TRANSIENT INFRARED SPECTROSCOPY OF CHARGE TRANSPORT IN EMERGING PHOTOVOLTAIC MATERIALS A Dissertation in Chemistry by Kwang Seob Jeong 2013 Kwang Seob Jeong Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2013

Upload: others

Post on 06-Dec-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

Department of Chemistry

TRANSIENT INFRARED SPECTROSCOPY OF CHARGE TRANSPORT IN

EMERGING PHOTOVOLTAIC MATERIALS

A Dissertation in

Chemistry

by

Kwang Seob Jeong

2013 Kwang Seob Jeong

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

December 2013

ii

The dissertation of Kwang Seob Jeong was reviewed and approved* by the following:

John B. Asbury

Associate Professor of Chemistry

Dissertation Advisor

Chair of Committee

David L. Allara

Distinguished Professor of Chemistry and Professor of Polymer Science

Mark Maroncelli

Professor of Chemistry

JianXu

Associate Professor of Engineering Science and Mechanics

Barbara J. Garrison

Shapiro Professor of Chemistry

Head of the Department of Chemistry

*Signatures are on file in the Graduate School

iii

ABSTRACT

Colloidal quantum dot (CQD) photovoltaic and organic photovoltaic (OPV) materials are

promising alternative light absorbers for solar cells. Both CQD photovoltaics and OPVs can be

fabricated on flexible substrates using low-cost solution cast fabrication methods at room

temperature. Although intense research has been done for the last two decades in both materials,

photophysical events underlying the device performance remain unclear. Here, the origin of the

charge transport state in PbS CQD solids was explored and identified. The charge transport state

was investigated using various optical and electrical methods: ultrafast transient infrared

spectroscopy (UFIR), microsecond transient infrared spectroscopy (TRIR), steady state

absorption spectroscopy, steady state photoluminescence emission spectroscopy, temperature

dependent TRIR, temperature dependent transient photoconductivity and temperature dependent

transient short-circuit current measurements. Furthermore, it was found that the mobility-lifetime

product, which is dependent on the surface passivation strategy, significantly influences the

device performance in CQD solar cells.

Additionally, it was examined how the dielectric permittivity influences the photophysics

in organic photovoltaic materials in conjunction with device performance. The experiments

revealed that the increase of dielectric permittivity leads to enhancement of the mobility-lifetime

product. For efficient conversion of excitons into charge carriers, it was suggested that high

surface area between electron donor and acceptor materials is necessary. The findings provide

better understanding of the fundamental properties of CQD and OPV materials and suggest

pathways to improve the efficiency of solar cell based on these materials.

iv

TABLE OF CONTENTS

List of Figures......... ................................................................................................................. vi

List of Tables ........................................................................................................................... xi

List of Abbreviations.................................................................................................................xii

Acknowledgements .................................................................................................................. xiii

Chapter 1 Introduction ............................................................................................................. 1

1.1 Motivation .................................................................................................................. 1 1.2 Basics of Solar Cells .................................................................................................. 2 1.3 Reference ................................................................................................................... 9

Chapter 2 Experimental methods ............................................................................................. 10

2.1 Sample preparation .................................................................................................... 10 2.2 Ultrafast transient infrared spectroscopy (UF-IR, Figure 2-1) .................................. 11 2.3 Microsecond transient infrared spectroscopy (TRIR, Figure 2-2) ............................. 12 2.4 Temperature dependent microsecond transient infrared spectroscopy ...................... 12 2.5 Organic photovoltaic material film ............................................................................ 13 2.6 Dispersive infrared spectroscopy ............................................................................... 13 2.7 Dielectric spectroscopy .............................................................................................. 13 2.8 Steady-state UV/Vis/NIR absorption spectroscopy ................................................... 14 2.9 Time Correlated Single Photon Counting (TCSPC) .................................................. 15 2.10 Steady-state photoluminescence spectroscopy......................................................... 16 2.11 Reference ................................................................................................................. 20

Chapter 3 Origin of sub-gap state and its role in charge transport in lead sulfide colloidal

quantum dot solid ............................................................................................................. 21

3.1 Introduction ................................................................................................................ 22 3.2 Result and Discussion ................................................................................................ 27 3.3 Conclusion ................................................................................................................. 34 3.4 Reference ................................................................................................................... 55

Chapter 4 Enhancement of mobility-lifetime product in PbS colloidal quantum dot solid. .... 58

4.1 Introduction ................................................................................................................ 59 4.2 Results and Discussion ............................................................................................... 60 4.3 Conclusion ................................................................................................................. 65 4.5 Reference ................................................................................................................... 76

v

Chapter 5 ATOMIC LIGAND PASSIVATED LEAD SULFIDE COLLOIDAL

QUANTUM DOT SOLID ............................................................................................... 78

5.1 Introduction ................................................................................................................ 79 5.2 Result and discussion ................................................................................................. 80 5.3 Conclusion ................................................................................................................. 86 5.4 Reference ................................................................................................................... 99

Chapter 6 Improvement of charge transport in organic photovoltaic materials by tuning

dielectric properties .......................................................................................................... 101

6.1 Introduction ................................................................................................................ 101 6.2 Results and Discussion ............................................................................................... 103 6.3 Conclusion ................................................................................................................. 108 6.4 Reference ................................................................................................................... 118

Chapter 7 Conclusion ............................................................................................................... 121

7.1 Summary .................................................................................................................... 121 7.2 Future direction .......................................................................................................... 122

7.2.1 Bias controlled transient two dimensional infrared spectroscopy

(BT2DIR) ......................................................................................................... 122 7.2.2 Time-resolved photoluminescence spectroscopy (TRPL) ...................................... 125 7.3Reference .................................................................................................................... 130

Appendix .................................................................................................................................. 131

Extension power law fit function ..................................................................................... 131

vi

LIST OF FIGURES

Figure 1-1. Schematic illustration of a p-n junction. ............................................................... 6

Figure 1-2. Forward (IF) and reverse (IR) bias in a p-n junction. ............................................. 7

Figure 1-3. J-V characteristic of a solar cell in the dark and under illumination ..................... 8

Figure 2-1. Ligand structure .................................................................................................... 17

Figure 2-2. Ultrafast transient infrared spectroscopy (UFIR) .................................................. 18

Figure 2-3. Microsecond transient infrared spectroscopy (TRIR) ........................................... 19

Figure 3-1. Electronic energy states of atom, semiconducting quantum dot and bulk

materials are illustrated. ................................................................................................... 35

Figure 3-2.UFIRdecaysof PbS CQD films. UFIR decays measured near the maximum of

mid-IR electronic transition (0.24 eV) in ligand exchanged PbS CQD solids. The

kinetic decays are normalized to highlight the ligand-dependent decay rates. The

inset shows the actual UFIR intensities of kinetic decays on logarithmic time scale. ..... 36

Figure 3-3.Mid-IR probe frequency dependent UFIR kinetic decays. UFIR rising curves

were measured at 0.24 and 0.33 eV. The rise times follows the instrument response

functions at both probe energies. ..................................................................................... 37

Figure 3-4. Temperature dependent UFIR decays of hybrid capped PbS CQD film. .............. 38

Figure 3-5. Inverse correlation of UFIR intensity and electron mobility of various ligand

capped PbS CQD films. ................................................................................................... 39

Figure 3-6. Transport-limited Auger recombination ................................................................ 40

Figure 3-7.The normalized UFIR kinetic decays measured at 0.33 eV following 1.55 eV

excitation represents that the UFIR decay is sensitive to pump excitation density. ........ 41

Figure 3-8. Mid-IR absorption and PL emission spectra of four different size

dots.ΔsindicatesStokes-shift energy which is a energy difference between the

excitonic absorption peak and the excitonic PL emission peak. ...................................... 42

Figure 3-9. Mid-IR absorption spectra for four different size hybrid capped PbS CQD

films. Bottom panels represent second derivation applied mid-IR absorption spectra

providing three common excitonic peaks (red curves). ................................................... 43

Figure 3-10. TRIR spectra measured 0.5 µs after photoexcitation for four different size

hybrid capped PbS CQD films. ........................................................................................ 44

Figure 3-11. Comparison of mid-IR transition energies (blue squares) to the sum of

Stokes-shift energy and 1Se-1Pe transition energy. The mid-IR transition energy

vii

quantitatively matches to the sum of Stokes-shift energy and 1Se-1Pe transition

energy. .............................................................................................................................. 45

Figure 3-12. Ligand dependent photoluminescence spectra of PbS CQD films. HDT >

EDT > MPA >Br . ............................................................................................................ 46

Figure 3-13.Comparison PL intensity to TRIR intensity. The quantitative agreement

between PL intensity and TRIR intensity indicates that the same populations of

photoexcited electrons are responsible for both mid-IR and PL transitions. ................... 47

Figure 3-14.Anti-correlation between UFIR and TRIR intensity with various ligand

passivation strategy. ......................................................................................................... 48

Figure 3-15.Temperature dependent TRIR provides activation energy of charge transport.

Decay rate of TRIR in logarithm scale is plotted as a function of time. From the

Arrhenius plots, the obtained activation energies are 8, 47 and 16 meV for 0.93, 1.30

and 1.46 eV bandgap energy hybrid capped PbS CQD film. ........................................... 49

Figure 3-16. Energy diagram of PbS CQD solid. .................................................................... 50

Figure 3-17(A) Origin of sub-gap transport state.Band structure diagram is plotted on the

basis of density functional calculations indicating distinct states of different

symmetry in the 1S state manifolds. Cartoons in the left side shows 1Se even (blue

top), 1Sh odd (red bottom), 1Se odd (blue top) and 1Sh even (red bottom). (B)

Localized trap states at surface of PbS CQD. .................................................................. 51

Figure 3-18. Reduction of photovoltage by Stokes shift increase. .......................................... 52

Figure 3-19.Transport activation energy related to polydispersity of CQDs. The half

widths at half maximum of the first exciton peaks of 0.93 and 1.46 eV bandgap

energy PbS CQD are compared.. ..................................................................................... 53

Figure 4-1. Molecular structure of ligands: Oleic acid (OA), Ethanedithiol (EDT) and 3-

mercaptopropionic acid (MPA)........................................................................................ 66

Figure 4-2. Infrared spectra of oleic acid (OA), 3-mercaptopropionic acid (MPA),

ethanedithoil (EDT), propanedithiol (PrDT), butanedithiol (BDT), pentanedithiol

(PnDT), hexanedithiol (HDT), 1,2-benzenedithiol (1,2-BzDT) and 1,4-

benzenedithiol (1,4-BzDT) capped PbS CQD films. ....................................................... 67

Figure 4-3. Scanning transmission electron microscopy (STEM) images of EDT capped

PbS CQD and MPA capped PbS CQD films. Average interparticle distance for the

EDT capped film and the MPA capped film are 2.1 nm and 1.8 nm, respectively. ......... 68

Figure 4-4. Microsecond transient infrared (TRIR) spectra of EDT capped PbS film

(green) and MPA capped PbS film (blue). The narrow vibrational features result

from ligands that are electronically perturbed by charges carriers. The broad

absorption features corresponds to subgap-to-1Pe transition. .......................................... 69

viii

Figure 4-5. Microsecond transient infrared spectra of various ligand capped PbS films.

The microsecond transient infrared spectra is strongly dependent upon the ligand

passivation. The maximum transition energy shift is about 0.13 eV and the area of

the spectra is correlated to the length of ligand. ............................................................... 70

Figure 4-6.Excitonic absorption peaks of 3-mercaptopropionic acid (MPA), ethanedithoil

(EDT), propanedithiol (PrDT), butanedithiol (BDT), pentanedithiol (PnDT),

hexanedithiol (HDT), 1,2-benzenedithiol (1,2-BzDT) and 1,4-benzenedithiol (1,4-

BzDT) capped PbS CQD films. The bandgap energy is dependent on ligand

passivation and the maximum shift is ~0.07 eV. ............................................................. 71

Figure 4-7.Mid-IR transition and bandgap correlation data. The scattered dots of various

ligand capped PbS CQD film implies the mid-IR transition is not strongly dependent

on quantum confinement. ................................................................................................. 72

Figure 4-8. TRIR decays of the EDT capped film (green) and the MPA capped film

(blue). The MPA capped film has lower lifetime by a factor of 3 in comparison to

the EDT capped film. ....................................................................................................... 73

Figure 4-9. Id-Vg curves for minority carriers (electrons) in the EDT and MPA capped

PbS Films. The mobility was measured using the ion-gel field-effect transistor

geometry. The inset is the magnified Id-Vg curve for the EDT capped PbS film for

clarity. .............................................................................................................................. 74

Figure 5-1. Atomic ligand passivation. Cd2+

cations of Cd-TDPA passivate S2-

dangling

bonds of PbS CQD at S1 step and the OA is replaced by bromide using CTAB at S2

step. .................................................................................................................................. 87

Figure 5-2. Infrared spectra of PbS CQDs before and after ligand exchange. ......................... 88

Figure 5-3. Scanning transmission electron microscopy (STEM) images of oleic acid

capped PbS CQD (A), EDT capped PbS CQD (B), MPA capped PbS CQD (C) and

Br capped PbS CQDs (D). From the images, the interparticle distances of each

samples were measured: 2.1 nm of EDT capped PbS CQD; 1.8 nm MPA capped

PbS CQD; 1.0 nm Br capped PbS CQD. ......................................................................... 89

Figure 5-4. Near-infrared absorption spectra of oleic acid capped PbS CQD and

transmission spectra for Br, Cl and I capped PbS CQD film. .......................................... 90

Figure 5-5. Near-infrared absorption spectra of Br, MPA, 1,4-BzDT, BDT and EDT

capped PbS CQD films. As decreasing ligand size, bandgap peak becomes broaden

and red-shifted. ................................................................................................................. 91

Figure 5-6. Id-Vg (black) and C-Vg (red) curves of hybrid capped PbS CQD film. ................. 92

Figure 5-7. Charge Extraction via a Linearly Increasing Voltage (CELIV) measurement

of Br capped PbS CQD film. Dielectric constant of the Br capped PbS CQD film

was measured to be 25±2. ............................................................................................... 93

ix

Figure 5-8. Transient infrared spectra of Br, MPA and EDT capped PbS CQD films at 0.5

µs after photoexcitation with 532 nm pump pulse. .......................................................... 94

Figure 5-9. The average decay times of the trap-to-band transitions in PbS CQD films are

strongly correlated with the average mid-infrared transition energies ............................. 95

Figure 5-10.The mobilities of electrons in PbS CQD films are also strongly correlated

with the average mid-infrared transition energies. ........................................................... 96

Figure 5-11. Microsecond transient infrared spectra of halide capped PbS CQD film (Br,

Cl and I) ........................................................................................................................... 97

Figure 6-1.Structure of M-TQ1, Lithium Bis(Trifluoromethanesulfonyl)Imide (LiTFSI)

and Phenyl-C61-Butyric-Acid-Methyl Ester (PCBM). .................................................... 109

Figure 6-2. Real dielectric functions (A) and real impedance functions (B) of M-TQ1

(red), 0.010 Li+/O (black), 0.015 Li+/O (green) doped M-TQ1 films............................. 110

Figure 6-3. Source-drain current, ISD, versus gate voltage, Vg, measured in thin films of

M-TQ1 doped with various concentrations of LiTFSI. The off current at positive

gate voltage increases substantially with lithium ion doping. .......................................... 111

Figure 6-4. Charge recombination kinetics decays of 1:1 by mass of M-TQ1 PCBM

polymer blend doped with various concentrations of lithium ions. The transient

signal arises from the polaron absorption in the polymer phase measured at 1700 cm-

1, shaded region within the inset. The gray curve represents the instrument response

function (IRF) used to deconvolute the kinetic decays. The data indicate slower

charge recombination dynamics with increasing lithium ion concentration. ................... 112

Figure 6-5.(A)Li ion dependent mobility and bimolecular recombination lifetime of M-

TQ1 polymer film.(B)Both the hole mobility and recombination lifetime give rise to

large increases of the mobility-lifetime product. ............................................................. 113

Figure 6-6.The absorption spectrum of a film of Li-doped M-TQ1 is almost identical to

that of pure M-TQ1. ......................................................................................................... 114

Figure 6-7. Steady-state fluorescence spectra of the pure M-TQ1, 0.010 Li+/O doped M-

TQ1, M-TQ1/PCBM 1:1 blend by mass and 0.010 Li+/O doped M-TQ1/PCBM 1:1

blend by mass. .................................................................................................................. 115

Figure 6-8. Time correlated single photon counting decays of M-TQ1 films in the

presence and absence of lithium ions are displayed. Interestingly, the exciton

lifetime in M-TQ1 is not significantly affected by the presence of Li ions. Addition

of PCBM to the pure polymer film causes a modest decrease in the exciton lifetime. .... 116

Figure 6-9. Energy filtered TEM image of a 1:3by massM-TQ1 PCBM polymer blend.

The image shows a carbon map such that the lighter gray regions correspond to the

PCBM-rich phase and the dark gray correspond to the M-TQ1 domain. The image

represents that M-TQ1 and PCBM undergo macroscopic phase separation. ................... 117

x

Figure 7-1. H-NMR of dimercaptoamid (DMA) ..................................................................... 126

Figure 7-2. Linear infrared spectrum of DMA ........................................................................ 127

Figure 7-3. Broad infrared pump probe spectrum of DMA capped PbS CQD film. ............... 128

Figure 7-4. Device geometry of BT2DIR ................................................................................ 129

xi

LIST OF TABLES

Table 3-1.Activation energies for electron transport states in PbS CQD films. ...................... 54

Table 4-1. Electrical and optical parameters of EDT capped PbS CQD film and MPA

capped PbS CQD film. ..................................................................................................... 75

Table 5-1. X-ray photoelectron spectroscopy (XPS) analysis of elemental atomic

percentage ........................................................................................................................ 98

Table 5-2. Binding energy and dielectric constant correlation ................................................ 98

xii

List of Abbreviation

OA Oleic acid

EDT Ethanedithiol

MPA 3-mercaptopropionic acid

PrDT 1,3-propanedithiol

BDT 1,4-butanedithiol

PnDT 1,5-pentanedithiol

HDT 1,6-hexanedithiol

1,2-BzDT 1,2-benzenedithiol

1,4-BzDT 1,4-benzenedithiol

FET Field-effect-transistor

TFT Thin-film-transistor

TRIR Microsecond Transient Infrared Spectroscopy

UFIR Ultrafast Transient Infrared Spectroscopy

PL Photoluminescence spectroscopy

P3HT Regioregularpoly(3-hexylthiophene)

PCBM Phenyl-C61-butyric acid methyl ester

CQD Colloidal quantum dot

OPV Organic photovoltaic material

TRPL Time-resolved photoluminescence spectroscopy

BT2DIR Bias controlled transient two dimensional infrared spectroscopy

xiii

ACKNOWLEDGEMENTS

First and foremost, I would like to express my sincere and profound gratitude to my

advisor, Prof. John Asbury, for his inspirational guidance and encouragement. Especially, his

consideration and patience helped me to overcome a number of challenges both academically and

personally. I will never forget the prayers you did for me throughout the last five years. I am

blessed to have worked with you.

I also want to thank my committees Prof. Allara, Prof. Maroncelli and Prof. Xu for

insightful suggestions and collaborative works. Your advice and insights broadened my

academic interest and knowledge.

I would like to deeply thank the Asbury group members for fruitful discussion and

enjoyable graduate life. I also desire to express my sincere gratitude to the UP, Timothy, Prisca

and Aquila small groups of Young Kwang Korean Presbyterian Church. The many small group

gathering we had were very meaningful and encouraging.

I wholeheartedly offer my sincere gratitude to my wife, Jihye Kim. She is my best friend

as well as an amazing colleague. Without you, I would not have been able to start and complete

graduate school. It has been such a wonderful time to be with you everywhere and anytime. I

also give thanks to my daughter, Lydia, for all the special memories we had the last two years.

You always make us smile.

I owe a great deal of gratitude to my parents, parents-in-law, my brothers and sister-in-

law for supporting me in any form of cheers and prayers.

Lastly, my utmost gratitude goes to God, who created me and gave me salvation. Without

your guidance, I would not have stood at all. Thanks, God.

1

Chapter 1 Introduction

1.1 Motivation

Global energy consumption has dramatically increased during the last six decades along

with the increase of human population and the rapid economic development. Currently, 15 TW is

consumed to manage the human energy demand, and the global energy consumption is predicted

to be over 30 TW by 2050.1,2

While current energy source mostly relies on fossil fuels,

researchers have turned their attention to environmentally friendly energy sources due to

environmental concerns. Solar energy has been frequently highlighted due to its abundance and

cleanness. 120,000 TW of solar power irradiate the Earth, which is much larger than the human

energy consumption. Currently, silicon is the most commonly used photovoltaic material. Its

power conversion efficiency (PCE) is over 20%3 which is close to the theoretical limit of a single

PN junction solar cell (33.7 %), also known as the Shockley-Queisser limit.4 However, due to

high fabrication costs, there is a demand to replace the silicon solar cell with flexible photovoltaic

materials with lower-cost fabrication methods.

Solution-cast approaches enable solar cells to be deposited on various substrates at room

temperature, enabling low-cost manufacturing. Colloidal quantum dots (CQDs) and organic

photovoltaic (OPV) materials are promising light absorbers to meet the flexibility and low-cost

demands. Both CQD solar cells and OPV solar cells have accomplished dramatic improvement

with the highest power conversion efficiencies (PCE) to date being 7 %5 and 12 %, respectively.

6

However, the underlying principles of how these types of solar cells work remain unclear. In

particular, considering the nanoscale size of the photovoltaic materials, understanding at the

molecular level is critical. In this work, we investigate the fundamental photophysics of CQD

solar cells and OPV solar cells using optical and electrical methods.

2

1.2 Basics of Solar Cells

Solar cells, or photovoltaic devices, are composed of semiconducting materials and

electrodes. Upon photoexcitation above the bandgap of the semiconducting material, the solar

cell material produces charge carriers that flow into their respective electrodes. The performance

of the solar cell device is determined by three types of efficiencies: the photon absorption

efficiency, the photon-to-charge carrier conversion efficiency and the efficiency of charge carrier

transport to the electrodes.

A PN junction consisting of a p-type and an n-type semiconducting material is the

building block of most semiconductor based solar cells (Figure 1-1). The p-type material has

excess holes resulting from p-type doping while the n-type material has excess electrons resulting

from n-type doping at thermal equilibrium. Both p-type and n-type semiconducting materials

have an intrinsic bandgap energy corresponding to the energy difference between the valence

band and the conduction band. Therefore, if a photon with energy larger than the bandgap is

absorbed, electrons in the valence band are excited to the conduction band leaving a hole in the

valence band and producing an exciton, which is a bound electron-hole pair. Then the excitons

are dissociated to generate free electrons and holes.

When the p-type material and the n-type material are brought together, excess electrons

in the n-type material diffuse to the p-type material and the excess holes in the p-type material

diffuse to the n-type material, creating a depletion region. Simultaneously, charges are left in the

semiconducting material and generate an electric field. In other words, when the electrons in the

n-type materials diffuse to the p-type material due to the doping density difference between two

doped materials, positively charged atoms are left behind in n-type material. On the other hand,

negatively charged atoms are left once the holes in p-type material diffuse to the n-type material.

Hence, due to charged atoms in the depletion region, an intrinsic electric field is formed in the

3

direction from the n-type material towards the p-type material. This electric field is known as the

Built-in Electric Field. The energy difference between the quasi-Fermi level of the n-type (EFN)

and the quasi-Fermi level of the p-type (EFP) is the built-in-potential barrier determining the

photovoltage of the solar cell. The built-in-potential barrier is described as the following

equation.

𝑉𝑏𝑖 = 𝑘𝑇

𝑒ln

𝑁𝑎𝑁𝑑

𝑛𝑖2

where k, T, e, Na, Nd, ni is the Boltzmann constant, the temperature, the elementary charge, the

acceptor concentration, the donor concentration and the intrinsic carrier concentration,

respectively. Therefore, under ambient condition, the built-in-potential barrier is determined by

the doping density of each semiconducting materials and the intrinsic carrier concentration.

Upon photoexcitation, charge carriers are generated in the depletion region and are

transported through drift toward the respective electrodes by the built-in-potential in the depletion

region. The width of the depletion region is described as

𝑊 = 2Ɛ 𝑉𝑏𝑖 + 𝑉𝑅

𝑒 𝑁𝑎 + 𝑁𝑑𝑁𝑎𝑁𝑑

1/2

where Ɛ, Vbi, VR, e, Na, Nd is the dielectric constant of the semiconductor, the built-in-potential

barrier, the applied reverse-bias voltage, the elementary charge, the acceptor concentration, the

donor concentration, respectively. The width of the depletion region is determined by the doping

concentration of each semiconductor and built-in-potential barrier. The increase of doping

4

density results in the increase of the built-in-potential barrier, but the width of depletion region is

reduced.

Charge carriers are transported via diffusion in the quasi-neutral region where there is no

built-in electric field. Due to the absence of built-in electric field, transport in the quasi-neutral

region is less efficient than in the depletion region because diffusion is less efficient than drift.

Thus, a long diffusion length is critical for efficient charge collection at the electrodes. For

example, when the diffusion length is larger than the difference between the solar cell film

thickness and the depletion width, more photogenerated charges can reach the electrodes. In

contrast, when the diffusion length is smaller than the difference between the solar cell film

thickness and the depletion width, charge carriers are more likely to recombine at recombination

centers in the quasi-neutral region. In this case, charge carriers are not efficiently collected

thereby limiting the power conversion efficiency of the solar cell device. Therefore, intensive

research has been done to enhance the diffusion length of charge carriers. The diffusion length is

described by the following equation.

𝑙𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 = 𝑘𝑇µτ

𝑒

where µ, τ, k, T, e is the charge carrier mobility, the charge carrier lifetime, Boltzmann constant,

temperature, elementary charge, respectively. As shown in the equation, the diffusion length is

related to the mobility and lifetime of charge carriers.

The solar cell efficiency, also referred to power conversion efficiency (PCE), is obtained

by measuring current density as a function of voltage under illumination. When positive voltage

is applied to the p-type material and negative voltage to n-doped material (Figure 1-2), the bias is

referred to a forward bias. In this case, the quasi-Fermi energy difference between the n-type and

the p-type is reduced since the direction of the applied external electric field is against that of the

intrinsic internal electric field in the depletion region.

5

In contrast, when a reverse bias is applied to the device which is the same direction with

the intrinsic internal electric field, additional voltage difference is added to the built-in-potential

barrier, producing a larger voltage drop and eventually reaching the open circuit voltage (Voc),

one of the important parameters determining solar cell PCE (Figure 1-3). The power conversion

efficiency (η) is described as the below equation.

𝜂(%) = 𝑉𝑜𝑐 𝐽𝑠𝑐𝐹𝐹

𝑃𝑖𝑛𝑐 × 100

where Voc, Jsc, FF, Pinc is the open-circuit voltage, the short-circuit current density, the fill factor

and the incident light intensity, respectively. In the absence of light, the current density remains

zero at negative bias and starts rising when it crosses Voc. However, in the presence of light, the

current density shows negative value at negative voltage and generates an area at the fourth

quadrant. The largest rectangular area corresponds to the maximum power output and the

intercepts of the curve are Jsc and Voc. The solar cell efficiency can be improved by increasing

Voc and/or Jsc. Voc is related to the photovoltage and the Jsc is involved with the charge

collection. Chapter 3 is related to the photovoltage of solar cell device and offers a guideline how

to improve the photovoltage in device by finding the energetic position of the charge transport

state in PbS colloidal quantum dot solar cell. Chapter 4, 5 and 6 involve studies on the

enhancement of diffusion length by changing the fundamental property of colloidal quantum dot

and organic photovoltaic materials. The diffusion length, which is determined by mobility-

lifetime product, is closely related to the charge harvesting determining the device performance.

Therefore, by varying sample condition such as ligand passivation and dielectric permittivity, the

mobility-lifetime product shows improvement, giving rise to the device performance

improvement.

6

Figure 1-1. Schematic illustration of a p-n junction.

7

Figure 1-2. Forward (IF) and reverse (IR) bias in a p-n junction.

8

Figure 1-3. J-V characteristic of a solar cell in the dark and under illumination

9

1.3 Reference

1Graetzel, M.; Janssen, R. A.; Mitzi, A. B.; Sargent, E. H. Materials Interface Engineering for

Solution-processed Photovoltaics. Nature, 2012, 488, 304-312.

2Wadia, C.; Alivisatos, A. P.; Kammen, D. M. Materials Availability Expands the Opportunity

for Large-Scale Photovoltaics Deployment. Environ. Sci. Technol. 2009, 43, 2072-2077.

3Green, M. A.; Emery, K.; Hishikawa, Y.; Warta, W. Solar Cell Efficiency Tables. Prog.

Photovoltaics,2010, 18,144–150.

4 Shockley, W.; Queisser, H. J. Detailed Balance Limit of Efficiency of p-n Junction Solar Cells.

J. Appl. Phys.,1961, 32, 510-519.

5Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, R.; Debnath, R.; Levina, L.;

Rollny, L. R.; Carey, G. H.; Fischer, A.; Kemp, K. W.; Kramer, I. J.; Ning, Z.; Labelle, A. J.; Wei

Chou, K.; Amassian, A.; Sargent, E. H. Hybrid Passivated Colloidal Quantum Dot Solids Nature

Nanotech., 2012, 7, 577-582.

6 www.heliatek.com

10

Chapter 2 Experimental methods

2.1 Sample preparation

PbS colloidal quantum dot films were fabricated on various substrates for the

measurement of UF-IR, TRIR, FET, Dispersive-IR, UV/Vis Absorption, TEM and PCE. The film

fabrication methods are almost identical although the sample substrates are different. Oleic acid

capped PbS CQDs dispersed in octane solution (50 mg/mL) were spun on a substrate at 2500

rpm. Secondly, shorter length ligand solutions such as 10% v/v 3-mercaptopropionic acid (MPA)

methanol solution were deposited on the PbS CQD film and the oleic acid surfactants were

replaced by the shorter organic ligands. Lastly, the ligand exchanged PbS CQD films were rinsed

with methanol and octane. The three steps were iterated for 10 times producing ~250 nm thick

films and this method is referred to as the layer-by-layer method (LBL). Acetonitrile was used as

the solvent for thiol solutions instead of methanol. Ethanedithiol (EDT), propanedithiol (PrDT),

butanedithiol (BDT), pentanedithiol (PnDT), hexanedithiol (HDT), 1,2-benzenedithiol (1,2-

BzDT) and 1,4-benzendithiol (1,4-BzDT) methanol solutions were used (Figure 2-1)

Halide ligand passivation strategy is slightly different from the thiol ligand film

fabrication method. Cd-treated PbS CQDs in octane solution (50 mg/mL) were dispersed on a

substrate and spun for 15 s at 2500 rpm. Cetyltrimethylammonium bromide (CTAB) in methanol

solution (10 mg/mL) was used as the bromide (Br-) source. Due to the slow ligand exchange rate

of CTAB, the CTAB solution was dispersed for 1 min and spun for 3 s. The ligand exchanged

PbS CQD film was rinsed with methanol and spun dry for 10 s. The ligand exchange and rinsing

processes were iterated for 10 times. The layer-by-layer method produced smooth and shiny

bromide capped PbS CQD films.

11

Film fabrication methods are identical for hexadecyltrimethylammonium chloride

(HTAC) and tetrabutylammonium iodide (TBAI) capped PbS CQD film preparation except for

the concentration of the ligand solution; 8.8 mg/mL for HTAC methanol solution and 10.1

mg/mL for TABI methanol solution. The halide ligand passivated PbS CQD films were made at

ambient condition with low relative moisture (<30%). CaF2 and TiO2 were used as substrates for

the microsecond transient infrared spectroscopy and the electrical measurement, respectively.

The hybrid capped PbS CQD film is prepared by following the LBL method. The oleic

acid capped PbS CQD solution (50 mg/mL) was spin-cast at 2500 rpm and 1% v/v 3-

mercaptopropionic acid methanol solution was used for ligand exchange. To note, the inorganic

passivation was performed through CdCl2 pretreatment during PbS CQD synthesis. The CdCl2

pretreatment covers most of Pb2+

empty orbitals at the surface with chloride and the nonbonding

electron orbitals of sulfur with cadmium ions.

2.2 Ultrafast transient infrared spectroscopy (UF-IR, Figure 2-2)

An ultrafast Ti:sapphire laser generates 800 nm wavelength centered pulses with 100 fs

duration at 1 kHz repetition rate and the beam is split by a beam splitter in front of the optical

parametric amplifier (OPA). One beam is used as a pump to photo-excite the CQD films in a

cryostat (Janis) under vacuum, and the other is tuned to mid-IR range (3-6µm) through an OPA

(Quantronix) to be used as a probe. Both pump and probe pulses overlap on the sample with 500

µm and 300 µm beam diameters, respectively. The mid-IR probe pulses reach a 64-element

mercury cadmium telluride (MCT) dual array detector (Infrared Associates, Inc.) to

simultaneously detect 32 probe frequencies through a monochromator (Horiba Jobin-Yvon). All

the instruments are purged with dry air. The experiments were carried out at room temperature

controlled by air handling equipment (Cleanpak).

12

2.3 Microsecond transient infrared spectroscopy (TRIR, Figure 2-3)

A second harmonic (532 nm) pulse generated by a Nd: YAG laser with 10 ns duration was

utilized for photo-excitation of the lead sulfide colloidal quantum dot film. Continuous-wave

infrared probe light emitted from a compact ceramic Globar Light Source (Newport Corp.) was

overlapped with the 532 nm excitation laser pulse of 500µJ/cm2 excitation density corresponding

to less than one photon per dot at the PbS CQD film sample. Both MPA and EDT capped PbS

CQD films were measured at the same excitation density and have similar optical absorptions at

532 nm. The PbS CQD film samples were deposited on a 1 inch protected Al mirror (Thorlabs)

and kept under vacuum at 10-2

Torr inside a vacuum cryostat (Janis). The infrared probe light

interacting with the PbS CQD film sample was reflected and dispersed by a 150 lines/mm grating

inside a monochromator (HORIBA JobinYvon). The stepwise grating guided the corresponding

wavelength infrared light into a 1x1 mm2 mercury cadmium telluride (MCT) single-element

detector. The detected signals were amplified by three orders of magnitude by a preamplifier

with band-pass frequency of 1.5 Hz to 1.0 MHz (Infrared systems Development) and digitized by

a 16-bit 10 MHz computer mounted analog-digital converter (National Instruments).

2.4 Temperature dependent microsecond transient infrared spectroscopy

The temperature dependent microsecond transient infrared spectroscopy was performed

by Jihye Kim in the Asbury group in the Chemistry Department at Penn State University. The

setup is identical to the methods described above, except that the sample was mounted inside a

cryostat (Janis, ST-100) and the temperature was controlled by a temperature controller (Lake

Shore).

13

2.5 Organic photovoltaic material film

Solutions of M-TQ1 and PCBM (1:3 by mass) were prepared in chlorobenzene with a total

concentration of 20 mg/mL and different concentrations of LiTFSI methanol solution were added

to the M-TQ1 and PCBM solution. The polymer blend solution was spun on 1 inch protected Al

mirrors (Thorlabs) at 500 rpm for 200 s under nitrogen atmosphere inside a glovebox.

2.6 Dispersive infrared spectroscopy

Steady state infrared spectroscopy is conducted by using home-built dispersive infrared

spectrometer and Fourier transform infrared spectrometer (Thermo Fisher Scientific, Inc.). The

dispersive infrared spectrometer utilizes a globar infrared ceramic element (6575, Newport)

hitting the sample amounted at a cold finger inside a cryostat (ST-100, Janis). The infrared light

interacting with a sample is dispersed at a grating (150 lines/mm, 5 µm blazed wavelength) inside

a monochromator (Horiba Jobin-Yvon, Inc.) through 0.5 cm size slit and 30 Hz optical chopper

(Newport Corp.). The grating mechanically moves stepwise and corresponding wavelength

infrared light is detected by a single element (1x1mm2) mercury cadmium telluride (MCT)

detector ( Infrared Associates Inc.) connected to a current mode preamplifier (Infrared Systems

Development Corp.). The signal intensity is amplified by about three orders of magnitude by the

preamplifier followed by analogue-to-digital converter (National Instruments Corp.).

2.7 Dielectric spectroscopy

Dielectric spectroscopy was carried by using an Agilent E4980A precision LCR meter in

the Wang group in the Department of Materials Science and Engineering at Penn State

14

University. The drop-casted sample film lies between two brass electrodes to be a sandwich type

cell (electrode-sample film-electrode) in the glovebox. The LCR meter measured the dielectric

constant and impedance functions in the AC field in a frequency range of 2 Hz - 2 MHz and N2

atmosphere was kept during the measurement.

2.8 Steady-state UV/Vis/NIR absorption spectroscopy

The UV/Vis/NIR spectra were obtained with a Beckman DU-520 UV/Vis/NIR

Spectrophotometer (Beckman Coulter Inc.). The UV/Vis/NIR spectrometer comprises of a

tungsten element for visible light and a deuterium element for UV light. A 1200 lines/mm

grating spatially separates the UV/Vis/NIR light and a silicon photodiode detects the

corresponding light. The sample materials are spin coated on 1 inch diameter calcium floride

(CaF2) windows (Red optronics) and a blank calcium floride window was used for background

offset.

*The following experiment were carried out by collaborators.

1) The Sargent group in the Department of Electrical and Computer Engineering at the University

of Toronto in Canada.

Ion-gel Field Effect Transistor (FET) 1

Current-Voltage (I-V) measurement1

Capacitance-Voltage (C-V) measurement1

Steady-state linear infrared spectroscopy1

15

Scanning transmission electron microscopy (STEM)1

X-ray diffraction analysis1

X-ray Photoelectron Spectroscopy1

Carrier extraction by linear increasing voltage (CELIV)1

Temperature-dependent transient photoconductivity measurement2

Temperature-dependent short-circuit current measurement2

2) The Maroncelli group in the Department of Chemistry at the Pennsylvania State

University

2.9 Time Correlated Single Photon Counting (TCSPC)

The time correlated single photon counting (TCSPC) experiment is carried out by Dr.

Minako Kondo in Dr. Mark Maroncelli’s group. A 400 nm laser pulse, generated by the second

harmonic generation of the cavity damped Ti: Sapphire Laser (Coherent Mira 900F) centered at

800 nm with 5.4 MHz repetition rate, was used for exciting sample and the photoluminescence

emission signal was collected by a single monochromator (ISA H10) around the emission

wavelength of 750 nm over a time interval out to 3 ns. The time resolution was determined to be

25 ps(FWHM) by measuring the scattering of the solution. The decay of PL emission as a

function of time was fitted using bi-exponential function convolved with the instrumental

response function. The decay was measured using front-face detection where-in the excitation

light was focusing on the front of the sample surface and the signal was collected from the same

region of the excitation with an angle of ~30⁰ selected to minimize the scattered and reflected

excitation light.

16

2.10 Steady-state photoluminescence spectroscopy

The steady-state visible photoluminescence emission spectra were measured by using the

PL spectrometer (SpexFluorolog 212) in the Maroncelli group. The sample is photo-excited at

400 nm and the fluorescence emission is collected from 350 nm to 1000 nm.

3) The Gomez group in the Department of Chemical Engineering at the Pennsylvania State

University

Thin Film Transistor (TFT)3

Power conversion efficiency (PCE)4

4) The Xu group in department of Engineering Science and Mechanics at the Pennsylvania State

University.

Steady-state photoluminescence emission spectroscopy

Steady-state infrared (IR)photoluminescence (PL) emission measurement was carried out

using a home-built PL spectrometer in the Xu group in the department of engineering science and

mechanics. The quantum dot samples on CaF2 window were excited at 650 nm by cw

semiconductor diode laser (10mW) and PL emission was collected at NIR PMT module (950-

1700 nm, Hamamatsu) through a monochromator (Acton Research Corporation). The 650 nm cw

laser power was 10 mW and the beam diameter was 3 mm, corresponding to ~1 photon per dot

excitation density.

17

Figure 2-1. Ligand structure

18

Figure 2-2. Ultrafast transient infrared spectroscopy (UFIR)

19

Figure 2-3. Microsecond transient infrared spectroscopy (TRIR)

20

2.11 Reference

1Tang,J; Kemp,K. W.; Hoogland, S.; Jeong, K.S.; Liu, H.; Levina, L.; Furukawa, M.; Wang, X.;

Debnath, R.; Cha, D.; Chou, K.W.; Fischer, A.; Amassian, A.; Asbury, J.B.; Sargent, E. H.

Colloidal-quantum-dot photovoltaics using atomic-ligand passivation Nature Mater.2011, 10,

765–771

2 Kemp, K. W.

ǂ; Jeong, K. S.

ǂ; Kim, J.; Voznyy, O.; Hoogland, S.; Thon, S. M.; Ip, A. H.;

Stewart, R. J.; Sargent, E. H.; Asbury, J. B. Dark State Mediate Photocarrier Transport in

Colloidal Quantum Dot Solid Submitted.

3Vakhshouri, K.; Gomez, E. D. Effect of Crystallization Kinetics on Microstructure and Charge

Transport of PolythiophenesMacromol. Rapid Commun.2012, 33, 2133-2137.

4Guo, C.; Lin, Y.; Witman, M. D.; Simth, K. A.; Wang, C.; Hexemer, A.; Strzalka, J.; Gomez, E.

D.; Verduzco, R. Conjugated Block Copolymer Photovoltaics with near 3 % Efficiency through

Microphase Separation Nano Lett.2013, 13, 2957-2963.

21

Chapter 3 Origin of sub-gap state and its role in charge transport in lead sulfide colloidal

quantum dot solid

This chapter was written on the basis of the following manuscript.

Kyle W. Kemp, Kwang S. Jeong, Jihye Kim, OleksandrVoznyy, Sjoerd Hoogland,

Susanna M. Thon, Alex H. Ip, Robert J. Stewart, Edward H. Sargent and John B. Asbury

Stokes Shifted Sub-gap State Mediate Photocarrier Transport in Colloidal Quantum Dot

Solids.

Submitted to Nature Communications.

*The work has been performed through collaboration with Dr. Edward H. Sargent's group in

department of electrical and computer engineering at the University of Toronto in Canada. The

electrical measurements (photoconductivity, short-circuit measurements), photoluminescence of

PbS CQD solution, and density functional theory computational studies were performed by the

Sargent group.

22

3.1 Introduction

Colloidal quantum dots (CQDs) provide unique advantages as photovoltaic materials

because of their size-tunable bandgap.1 The bandgap tunability through varying the nanoparticle

size enables us to utilize most of the solar spectrum including the near infrared range that contains

half of the energy from the sun.2 The tunable bandgap originates from the quantum confinement

of the semiconducting nanoparticle. Figure 3-1 represents the electronic states of a molecule, a

nanocrystal and a bulk material. As the bulk material shrinks its size to nanometer scale, the

continuous bandgap energy states of the material become discrete, similar to molecular electronic

energy states.

The quantum confinement is attributed to the natural length scale of the electron and hole

(Bohr radius) of a nanoparticle being comparable or longer than the physical quantum dot size.

The electronic states of the nanoparticle are analogous to the states in the hydrogen-like model.

Thus the electronic states are named as S, P, D..states. The Bohr radius in a semiconductor is

defined in terms of the Bohr radius of a hydrogen atom, a0,

𝑎𝐵 = Ɛ𝑚

𝑚∗𝑎0

where Ɛ is the dielectric constant of the bulk material, m* is the mass of the atomic particle, and

m is the rest mass of the electron. Due to Coulombic attraction between the electron and hole, the

exciton is regarded as a hydrogen-like bound state. Depending on the size of Bohr radius relative

to the size of nanoparticle, quantum confinement can be classified into three categories.

Strong confinement appears when the quantum dot radius is much smaller than electron

Bohr radius (ae), hole Bohr radius (ah) and exciton Bohr radius (aexc). Lead sulfide (PbS, aexc=

18nm) and lead selenide (PbSe, aexc= 20nm) quantum dots possess strong confinement in general

due to their relatively large Bohr radius. When the nanoparticle radius is smaller than aexc, but

larger than both ae, ah, this is referred to weak confinement regime. Lastly, if the nanoparticle

23

size lies between ae and ah or vice versa and smaller than aexc, an intermediate confinement regime

is observed.3

The bandgap energy of the strongly confined quantum dot was theoretically modeled by

L. Brus in 1984 using the effective mass approximation.1 The model well describes how the

bandgap of the nanoparticle depends on its radius,

𝐸𝑔𝑁𝑃 = 𝐸𝑔 +

ħ2𝜋2

2𝑅2 1

𝑚𝑒𝑓𝑓𝑒 +

1

𝑚𝑒𝑓𝑓𝑕 −

1.8𝑒2

Ɛ𝑅

where, Eg, ħ, R,𝑚𝑒𝑓𝑓𝑒 ,𝑚𝑒𝑓𝑓

𝑕 ,e, Ɛ are the bandgap of the bulk material, the reduced Planck

constant, the nanoparticle radius, the effective mass of electron, the effective mass of hole, the

elementary charge and dielectric constant, respectively. This equation starts from the bandgap of

the bulk material, Eg, which is smaller than that of a semiconducting nanoparticle. The second

term corresponds to the quantum confinement energy which is inversely proportional to the

square of the nanoparticle size, and the effective mass of electron and holes. The third term is

involved with Coulomb attraction of the electron and hole pair. This model describes the

bandgap energy of a strongly confined semiconducting nanoparticle.

The effective-mass approximation is useful for estimating the bandgap of a nanoparticle.

However, it does not offer a means of explaining sub-gap states that are placed between the

valence band and the conduction band. In bulk semiconducting material, the perfect periodicity

of the idealized single-crystal lattice is abruptly cut at the surface of material when the

semiconductor is sharply terminated. The abrupt termination of the periodicity of the potential

function gives rise to generation of allowed electronic energy states within the energy bandgap

where no transitions were formally allowed due to the absence of electronic states. Specifically,

at the surface, when assuming that the sp3 hybrid tetrahedral orbitals were used for bonding

between atoms, one of the sp3 orbital is left as a nonbonding orbital without overlapping with the

orbitals of adjacent atom. The nonbonding orbital does not produce any symmetric and anti-

24

symmetric energy splitting as in the case of overlapped orbitals, leading to the nonbonding orbital

being placed between the lower energy level of bonding molecular orbitals and the higher energy

level of antibonding molecular orbitals. The nonbonding orbital is referred to a dangling bond.

Furthermore, the presence of dangling bonds cause reconstruction of the surface structure to

stabilize the surface energy. The effective-mass approximation is not able to properly describe

this surface reconstruction.4 The lack of capability to include the electronic states of surface in

the effective-mass approximation is pronounced in semiconducting quantum dots. The

assumption of the effective mass approximation is that an electron is regarded as a free particle

owing to the large size of bulk material. However, the size of a quantum dot is less than 10 nm,

substantially smaller than that of the bulk material. Therefore, it is not proper assumption to fully

account for the electronic structure of quantum dot which has small size and large surface area.

The tight-binding approximation, alternatively, is based on the linear combination of

atomic orbitals (LCAO) principle applied to solid state materials. The approximation was

introduced to take account properties of nanocrystal surfaces.56

The tight-binding approximation

utilizes the atomic orbitals as base sets and calculates coordination between frontier orbitals of

atoms. Therefore, this approximation is able to include the nonbonding surface orbitals,

distinguishing the potential from that of core atoms of which orbitals have all possible

coordination completed. The tight-binding approximation provides the theoretical background of

the sub-gap electronic states of quantum dots which have been observed by photoluminescence

(PL) emission and electrical measurement.78

Nevertheless, due to its less accuracy of estimating

the bandgap energy, the tight binding approximation is complementary to the effective-mass

approximation.

Along with the two theoretical backgrounds for electronic states of colloidal quantum

dots, a number of experiments have been carried out for the last three decades. Steady state

visible absorption, photoluminescence emission, composition analysis using X-ray and optical

25

microscopy have been used to analyze quantum dot properties. In addition, time resolved

measurements such as transient absorption spectroscopy were also used to understand

photophysics in CQDs.9

Most studies on electronic states of colloidal quantum dots were conducted using the

colloidal phase of the quantum dots due to the simplicity to rule out electronic coupling between

quantum dots and the ease to conduct experiment. However, the studies for colloidal phase

quantum dots cannot directly be applied to solid state quantum dot films since the electronic

interaction between quantum dots becomes substantial in the solid phase, which is not an issue in

the study for solution phase colloidal quantum dots.

Recently, a dramatic improvement in quantum dot solar cell performance was reported,

where the power conversion efficiency of PbS colloidal quantum dot (CQD) solar cells achieved

7% in 2012.10

Surprisingly, it took only a few years to increase the power conversion efficiency

to 7% from 1%. While the device performance has developed at a very fast rate, understanding

the electronic structure of solid state colloidal quantum dot films has not kept pace with device

performance. Consequently, there is a need to explore the fundamental photophysics of the solid

state CQD photovoltaic materials. There has been intense research to understand charge transport

and recombination processes in semiconducting nanoparticle. Nevertheless, reported

mechanisms underlying the charge transport and recombination processes are still controversial.

Our work using ultrafast and microsecond transient infrared spectroscopy makes it

possible to probe mid-IR electronic transitions of photocarriers over seven orders of magnitude in

time.11,12,13,14,15

By scrutinizing the carrier dynamics, the charge transport and recombination

processes in PbS CQD solid film were investigated. It was found that a sub-gap state lies below

the 1Se state separated by the Stokes-shift energy which is the energy difference between the

bandgap absorption and emission spectra. In combination with temperature dependent studies

from transient absorption kinetic decays, transient photoconductivity and short-circuit current

26

measurements, we discovered unprecedented findings that the Stokes-shifted state is the charge

transport state. We believe that the discovery will contribute to the colloidal quantum dot field in

both understanding the electronic properties of PbS colloidal quantum dots and guiding direction

for device manufacture.

27

3.2 Result and Discussion

The ultrafast transient infrared spectroscopy (UFIR) was performed to examine the mid-

IR transition of various ligand capped PbS CQD films. A 100 fs duration 800 nm wavelength

pulse was used for bandgap excitation at 1 kHz repetition rate followed by a 200 fs duration mid-

IR probe pulse interacting with the photoexcited PbS CQD films. Surprisingly, the UFIR

absorption intensity and decay are sensitive to the ligand passivation strategy. The inset in Figure

3-2 represents the UFIR absorption decays of hybrid, 3-mercaptopropionic acid (MPA),

ethanedithiol (EDT) and hexanedithiol (HDT) capped PbS CQD films. Hybrid passivation

indicates the organic and inorganic mixed ligand passivation. The hybrid capped PbS CQD is

treated by CdCl2 during CQD synthesis to passivate Pb2+

and S2-

,and then MPA ligand exchange

was carried out via layer-by-layer method. The hybrid, MPA, EDT and HDT capped PbS CQD

films show 0.036, 0.027, 0.019, and 0.012 transient absorption (arbitrary units), respectively. The

ligand dependent transient absorption intensity does not fit the conventional electronic model of

colloidal quantum dots. The conventional model merely includes 1Se, 1Pe, 1De,etc. delocalized

electronic states (1Sh and 1Ph, etc. for valence band). The conventional view predicts the same

density of the charge carrier in 1Se states if the films absorb the same density of photons.

However, our result, which have been scaled by the density of absorbed photons, reveals that the

UFIR intensity is sensitive to ligand passivation. The UFIR intensity study indicates that

electrons are able to relax on ultrafast time scales into sub-gap states that do not exhibit near-IR

electronic transitions and that the density of these states depends on ligand treatment.

The ligand dependent UFIR intensity can be explained by comparing the trap density of

PbS CQD films treated with various ligands.16

The trap density of the PbS CQD film decreases

in the order of EDT, MPA and Br- capped films. Since the hybrid capped PbS CQD is mostly

passivated by chloride and additional post-passivation is treated by MPA, the trap density of

28

hybrid should be similar or perhaps even less than that of the Br- capped film. Based on the anti-

correlation between UFIR intensity and the trap density, it is revealed that the UFIR is dependent

upon the density of trap states, implying that charge carriers can partition between different types

of states. It was reported that charge trapping takes place at very fast time scales17

and is involved

with exciton dissociation, leading to reduction of PL quantum yield181920

. Our UFIR kinetic

traces also do not exhibit a fast decay that would be associated with trapping of electrons (Figure

3-3), indicating that the trapping process is faster than the UFIR instrument temporal resolution

(100 fs). The electron transfer time from PbS CQD to TiO2 was reported to be 6.4 ±0.4 fs,21

which means that the exciton dissociation occurs at least within a few fs. Furthermore, based on

the observation by Engel and co-worker, the electron cooling process completes within 100 fs,22

indicating charge carriers relax to the sub-gap state within 100 fs. Therefore, the measured UFIR

intensity in 1ps mostly represents photocarrier density at the sub-gap state.

Interestingly, the decay of the UFIR signal is also reliant on the ligand passivation. The

time constants were obtained from the extended power law fit function combined with two rising

fit functions (Details are described in the Appendix). The time constant for each films are 1.3 ns,

2.6 ns, 3.7 ns and 6.5 ns for hybrid, MPA, EDT and HDT capped PbS CQD film, respectively. In

order to examine the density of charge carrier closer to the time zero point (t=0), the UFIR decays

of 1.30 eV bandgap hybrid PbS CQD were collected in a temperature range from 133K to 293K

with 30K increment (Figure 3-4). The time constant noticeably increases with lowering

temperature while the intensity only changes by 7±4% from 133 K to 293K. This suggests that

the initial transient absorption intensity should not be significantly different from the intensity

measured at room temperature. From the temperature dependent study, we also learned from the

significant decay rate variation with temperature that this decay is related to charge transport

since the electron cooling, which is the other possible process, is not a non-equilibrium process

that is not expected to be sensitive to temperature. It is also found that the UFIR kinetic decay

29

rate is proportional to the mobility of PbS films (Figure 3-5). The electron mobility is determined

by the ligand passivation strategy: bromide capped film (4 x 10-2

cm2/V·s) > MPA acid capped

film (5.1x 10-3

cm2/V·s) > EDT capped film (2.4 x 10

-4cm

2/V·s).

2324 For all experiments, the

pump excitation density was kept under one photon per dot. Hence, the fast UFIR decay of the

hybrid capped PbS CQD film might result from bimolecular process such as transport-limited

Auger recombination.25

The transport-limited Auger recombination process is illustrated in

Figure 3-6. Traditional Auger recombination arises when one exciton in close proximity to

another donates its excitation energy to the second exciton resulting in loss of the first and double

excitation of the second. The doubly excited exciton is able to rapidly cool back to the lowest

energy excited state such that the net result of the process is loss of an exciton. Traditional Auger

recombination is important in quantum dots when there is a high probability that two excitons

will be formed in the same nanocrystal. However, in dense CQD solids, excitons and charge

carriers are mobile and are therefore able to migrate to neighboring nanocrystals. IN this

situation, excitons or charge carriers created in one nanocrystal can migrate to other nanocrystals

that also contain excitons or charge carriers. In this situation, bimolecular recombination

processes can occur even at excitation densities well below the threshold where the probability of

absorbing two excitons per nanocrystal is significant. In order to confirm whether a transport-

limited bimolecular recombination process such as Auger recombination occurs in the hybrid

capped PbS CQD films, we examined UFIR kinetic decays at various pump excitation densities

in the range of 0.11- 1.84 photon per nanoparticle (Figure 3-7). The excitation density dependent

UFIR decay measurements show that transport-limited bimolecular recombination process occurs

even at lower than one photon per dot excitation density.

In order to find the energy level of the sub-gap state, we measured steady-state

UV/Vis/NIR absorption, photoluminescence (PL) emission and TRIR spectra of different size

PbS CQDs. The samples were oleic acid capped PbS CQDs in octane with bandgaps of 0.93,

30

1.22, 1.30 and 1.46 eV. The energy differences between the near-IR (NIR) absorption peak and

the PL peak, referred to as the Stokes shifts (Figure 3-8), were obtained by comparing the two

spectra. Second derivative analysis was applied to the NIR absorption spectra to identify

quantum confined transitions (Figure 3-9). Three common quantum confined interband

(excitonic) transitions were found (red curvatures) in the four mid-IR absorption spectra.

Reported assignments of the second and the third absorption peaks are controversial. Some

assign the second interband transition as 1Sh-1Pe. This assignment is based on the hypothesis that

the broken symmetry arising from imperfect surface passivation and non-perfectly spherical

shape of quantum dots allows the symmetry forbidden transition (1Sh-1Pe). If this is the case, the

third peak is supposed to be 1Ph-1Pe. Others assigned the second interband transition to arise

from 1Ph-1Pe transition which is the same assignment we incorporated. The 1Se-1Pe energy

difference was calculated by half of the energy difference of the 1Sh-1Se and 1Ph-1Pe transitions.

Surprisingly, the sum of this calculated 1Se-1Pe splitting energy and the Stokes shift energy

quantitatively matches the transition energy of the TRIR spectra (Figure 3-10 and 3-11). This

assignment is consistent with the results of the Klimov group and the Jiang group.26,27,28

The

quantitative agreement corroborates the hypothesis that the energy level of the sub-gap state

probed by TRIR is below the 1Se optically bright state by Stokes-shift energy, and the transition

energy of TRIR corresponds to the Stokes-shifted sub-gap state to 1Pe transition energy.

Comparison of the ligand dependent intensities obtained from TRIR with those of N-IR

photoluminescence spectra further supports that the sub-gap state involved in the TRIR transition

peak is identical to the Stokes-shifted dark state (Figure 3-12 and 3-13). The ligand dependent PL

emission intensity quantitatively matches the TRIR intensities, implying that the Stokes-shifted

sub-gap state is the state involved with PL emission in PbS CQDs.

Interestingly, the UFIR intensity decreases in the order of hybrid, MPA, EDT and HDT,

which is in inverse order of intensity of TRIR for the same film (Figure 3-14). The most distinct

31

difference between UFIR and TRIR is time delay. The UFIR reflects charge carrier dynamics

from tens of femtosecond to nanosecond time scale and the TRIR represents the charge carrier

dynamics from a hundred nanosecond to millisecond timescale after photoexcitation.

The inverse correlated intensity of UFIR and TRIR can be understood by comparing the

UFIR decay rate to the mobility of carriers. Surprisingly, the trend in decreasing mobility of PbS

CQD films is in the order (halide > MPA > EDT). This trend is consistent with that of the UFIR

decay rate (halide > MPA >EDT). Considering that the UFIR decay is related to the transport-

limited bimolecular recombination process, the correlation between the UFIR kinetic decay and

mobility demonstrates that charge carriers are transported through sub-gap states. The UFIR

intensities indicate the density of photocarriers in the sub-gap states before the onset of charge

recombination while the TRIR kinetic trace indicates the density of charge carrier after a

significant fraction of carriers have been lost due to the recombination process.

The hybrid capped PbS CQD dots with different bandgap energies (0.93, 1.30 and 1.46

eV) were examined using temperature dependent TRIR spectroscopy to obtain charge transport

activation energy. Strikingly, it was found that the activation energy measured using TRIR

spectroscopy shows quantitative agreement within experimental precision with the activation

energy measured by the temperature dependent transient photoconductivity and short-circuit

current techniques (Figure 3-15 and Table 3-1). Furthermore, the activation energy is much less

than the Stokes shift energy, suggesting that charge transport is not mediated by the 1Se state. In

other words, the activation energy agreement indicates that the charge carrier transport process

occurs via the Stokes-shifted sub-gap state. The Table 3-1 tabulates the 1Sh-1Se (bandgap)

energy, the Stokes shift energy, the activation energies measured by temperature depended TRIR,

and the activation energies measured by temperature dependent transient photoconductivity and

Jsc. With all the observations, it is concluded that the Stokes-shifted sub-gap state is the charge

32

transport state which is also optically probed by the UFIR, the TRIR and the NIR PL. This

assignment is unprecedented in the CQD field (Figure 3-16).

Density functional theory (DFT) calculations were used to understand the origin of the

Stokes-shifted sub-gap states by the Sargent group at the University of Toronto. DFT

calculations indicates that the four-fold 1Se states are comprised of two different symmetry states

represented as even and odd symmetry in Figure 3-17(A).29

Under perfect surface passivation,

optical transitions between the nearest bandedge states are allowed. When defects are generated

on the Pb sub-lattice due to imperfect passivation, however, the electronic states of1Se are

distorted and the even and odd states are reordered. This is because the 1Se(even) state is sensitive

to the surface defects (Figure 3-17 (B)). The reordering of different symmetry states causes the

closest-lying electronic states to no longer be radiatively coupled to each other. This reordering

produces a large Stokes shift that corresponding to the energy between even and odd symmetry

1Se states.

The Stokes-shift is not desirable for solar cell device performance when it comes to

photovoltage. The photovoltage (related to the built-in-potential at the P-N junction) is

determined by the energy difference between p-type PbS CQD Fermi energy and n-type material

Fermi energy. (Figure 3-18) Therefore, reduction of the energy of the transport states near the

conduction band of p-type PbS CQD by the Stokes-shift energy results in a decrease o f the

maximum available photovoltage. Because the Stokes-shift is caused by imperfect bonding of

ligands to nanocrystal surfaces, removing charge traps and strain will help minimize the Stokes

shift energy and maximize the photovoltage of CQD PV devices.

Another approach for device performance improvement is to reduce the polydispersity of

colloidal quantum dots. Figure 3-19 represents the half-width-at-half-maximum (HWHM) of two

dots with different bandgap energies (0.95 eV and 1.49 eV bandgap). The HWHM of the

Gaussian fits indicates that the 0.93 eV PbS CQDs are highly monodisperse while the 1.46 eV

33

show broad size distribution. Interestingly, this trend is consistent with that of activation energies

measured by temperature dependent TRIR, transient photoconductivity and short-circuit current

measurement. The correlation between HWHM and charge transport activation energy suggests

that reduction of quantum dots size distribution may enhance photoconductivity by reducing

charge transport activation energy.

34

3.3 Conclusion

In conclusion, we revealed the origin of Stokes-shifted sub-gap state and its

characteristics in PbS CQD solid. While Stokes-shifted sub-gap states do not do not undergo

strong dipole-allowed optical transitions from 1Sh states in PbS CQDs, UFIR and TRIR

spectroscopies are capable of probing this state. By scrutinizing the mid-IR absorption spectra

and the photoluminescence emission spectra, we identified the energetic position of the Stokes-

shifted sub-gap state.

In combination with the temperature dependent studies of TRIR, transient

photoconductivity and the short-circuit current measurement, we discovered that the Stokes-

shifted sub-gap state is the charge transport state coupled to surface of quantum dot.

Based on the DFT calculation, the Stokes shift results from the electronic state exchange

within 1Se states due to presence of surface traps. For the future direction, reduction of surface

irregular structure and strain is required to remove the Stokes shift to increase the photovoltage.

Additionally, reduction of polydispersity is desired to reduce charge transport activation energy.

35

Figure 3-1. Electronic energy states of atom, semiconducting quantum dot and bulk materials are

illustrated.

36

Figure 3-2.UFIRdecaysof PbS CQD films. UFIR decays measured near the maximum of

mid-IR electronic transition (0.24 eV) in ligand exchanged PbS CQD solids. The kinetic decays

are normalized to highlight the ligand-dependent decay rates. The inset shows the actual UFIR

intensities of kinetic decays on logarithmic time scale.

37

Figure 3-3.Mid-IR probe frequency dependent UFIR kinetic decays. UFIR rising curves

were measured at 0.24 and 0.33 eV. The rise times follows the instrument response functions at

both probe energies.

38

Figure 3-4. Temperature dependent UFIR decays of hybrid capped PbS CQD film.

39

Figure 3-5. Inverse correlation of UFIR intensity and electron mobility of various ligand

capped PbS CQD films.

40

Figure 3-6. Transport-limited Auger recombination

41

Figure 3-7.The normalized UFIR kinetic decays measured at 0.33 eV following 1.55 eV

excitation represents that the UFIR decay is sensitive to pump excitation density.

42

Figure 3-8. Mid-IR absorption and PL emission spectra of four different size

dots.ΔsindicatesStokes-shift energy which is a energy difference between the excitonic absorption

peak and the excitonic PL emission peak.

43

Figure 3-9. Mid-IR absorption spectra for four different size hybrid capped PbS CQD

films. Bottom panels represent second derivation applied mid-IR absorption spectra providing

three common excitonic peaks (red curves).

44

Figure 3-10. TRIR spectra measured 0.5 µs after photoexcitation for four different size

hybrid capped PbS CQD films.

45

Figure 3-11. Comparison of mid-IR transition energies (blue squares) to the sum of

Stokes-shift energy and 1Se-1Pe transition energy. The mid-IR transition energy quantitatively

matches to the sum of Stokes-shift energy and 1Se-1Pe transition energy.

46

Figure 3-12. Ligand dependent photoluminescence spectra of PbS CQD films. HDT >

EDT > MPA >Br .

47

Figure 3-13.Comparison PL intensity to TRIR intensity. The quantitative agreement

between PL intensity and TRIR intensity indicates that the same populations of photoexcited

electrons are responsible for both mid-IR and PL transitions.

48

Figure 3-14.Anti-correlation between UFIR and TRIR intensity with various ligand

passivation strategy.

49

Figure 3-15.Temperature dependent TRIR provides activation energy of charge

transport. Decay rate of TRIR in logarithm scale is plotted as a function of time. From the

Arrhenius plots, the obtained activation energies are 8, 47 and 16 meV for 0.93, 1.30 and 1.46 eV

bandgap energy hybrid capped PbS CQD film.

50

Figure 3-16. Energy diagram of PbS CQD solid.

51

Figure 3-17(A) Origin of sub-gap transport state.Band structure diagram is plotted on the

basis of density functional calculations indicating distinct states of different symmetry in the 1S

state manifolds. Cartoons in the left side shows 1Se even (blue top), 1Sh odd (red bottom), 1Se odd

(blue top) and 1Sh even (red bottom). (B) Localized trap states at surface of PbS CQD.

A

B

52

Figure 3-18. Reduction of photovoltage by Stokes shift increase.

53

Figure 3-19.Transport activation energy related to polydispersity of CQDs. The half

widths at half maximum of the first exciton peaks of 0.93 and 1.46 eV bandgap energy PbS CQD

are compared..

54

Table 3-1.Activation energies for electron transport states in PbS CQD films.

1Sh-1Se (eV) Stokes Shift

(meV)

Temperature

Dependent TRIR

Ea (meV)

Temperature

Dependent

Ea (meV)

Temperature

Dependent Jsc

Ea (meV)

0.93 50 8±12 3±10 3±1

1.30 130 47±11 63±6 63±2

1.46 190 16±11 29±2 42±2

55

3.4 Reference

1Brus, L. Electronic Wave Functions in Semiconductor Clusters: Experiment and Theory. J.

Phys. Chem., 1986, 90, 2555-2560.

2 Sargent, E. H. Infrared Photovoltaics made by solution processing Nature. Photonics.2009, 3,

326-331.

3Klimov, V. I. Semiconductor and Metal Nanocrystals: Synthesis and Electronic and Optical

Properties; Marcel Dekker, INC.: New York, Basel, 2004.

4 Snoke, D. W. Solid State Physics: Essential Concepts; Addison-Wesley: San Francisco, 2009.

5 Slater, J. C.; Koster, G. F. Simplified LCAO Method for the Periodic Potential Problem. Phys.

Rev., 1954, 94, 1498-1524.

6Bryant, G. W.; Jaskolski, W. Surface Effects on Capped and Uncapped Nanocrystals. J. Phys.

Chem. B.2005, 109, 19650-19656.

7 Jones, M; Shun, S. L., Scholes, G. D. Quantitative Modeling of the role of surface traps in

CdSe/CdS/ZnSNanocrystal Photoluminescence Decay Dynamics. Proc. Natl. Acad. Sci., 2009,

106, 3011-3016.

8Nagpal, P.; Klimov, V. I. Role of Mid-gap States in Charge transport and Photoconductivity in

Semiconductor NanocrystalFilms Nat.Commun., 2011, 2, 486.

9Klimov, V. I.; Schwarz, Ch. J.; McBranch, D. W.; Leatherdale, C. A.; Bawendi, M. G. Ultrafast

Dynamics of Inter- and Intraband transition s in Semiconductor Nanocrystals: Implications for

Quantum-dot Lasers Phys. Rev. B, 1999, 60, 2177-2180.

10

Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, R.; Debnath, R.; Levina, L.;

Rollny, L. R.; Carey, G. H.; Fischer, A.; Kemp, K. W.; Kramer, I. J.; Ning, Z.; Labelle, A. J.; Wei

Chou, K.; Amassian, A.; Sargent, E. H. Hybrid Passivated Colloidal Quantum Dot Solids.

Nature Nanotech.,2012, 7, 577-582.

11

Shim, M.; Shilov, S. v.; Braiman, M. S.; Guyot-Sionnest, P. Long-Lived Delocalized Electron

States in Quantum Dots: A Step-Scan Fourier Transform Infrared Study. J. Phys. Chem. B, 2000,

104, 1494.

12

Wehrenberg, B. L.; Wang, C. J.; Guyot-Sionnest, P. Interband and Intrabad Optical Studies of

PbSe Colloidal Quantum Dots.J. Phys. Chem. B, 2002, 106, 10634.

13

Pandey, A.; Guyot-Sionnest, P. Intraband Spectroscopy and Band Offsets of Colloidal II-VI

Core/Shell Structures. J. Chem. Phys., 2007, 127, 104710.

56

14

Zhang, J.; Jiang, X. Confinement-Dependent Below-Gap State in PbS Quantum Dot Films

Probed by Continuous-Wave Photoninduced Absorption. J. Phys. Chem. B, 2008, 112, 9557.

15

Lewis, J. E.; Wu, S.; Jiang, J. Unconventional Gap State of Trapped Exciton in Lead Sulfide

Quantum Dots. Nanotechnology, 2010, 21, 455402.

16Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, R.; Debnath, R.; Levina, L.;

Rollny, L. R.; Carey, G. H.; Fischer, A.; Kemp, K. W.; Kramer, I. J.; Ning, Z.; Labelle, A. J.; Wei

Chou, K.; Amassian, A.; Sargent, E. H. Hybrid Passivated Colloidal Quantum Dot Solids.

Nature Nanotech., 2012, 7, 577-582.

17

McArthur, E. A.; Morris-Cohen, A. J.; Knowles, K. E.; Weiss, E. A. Charge Carrier Resolved

Relaxation of the First Excitonic State in CdSe Quantum Dots Probed with Near-Infrared

Transition Absorption Spectroscopy. J. Phys. Chem. B, 2010, 114, 14514-14520.

18Smith, A.M.; Duan, H.W.; Rhyner, M.N.; Ruan, G.; Nie, S.M.A Systematic Examination of

Surface Coatings on the Optical and Chemical Properties of Semiconductor Quantum Dots. Phys.

Chem. Chem. Phys.2006, 8, 3895-3903.

19Qu, L.H.; Peng, X.G. Control of Photoluminescence Properties of CdSeNanocrystals in growth.

J. Am. Chem. Soc., 2002, 124, 2049–2055.

20Talapin, D.V.; Rogach, A.L.; Kornowski, A.; Haase, M.; Weller, H. (2001) Highly luminescent

monodisperseCdSe and CdSe/ZnSnanocrystals synthesized in a hexadecylaminetrioctylphosphine

oxide-trioctylphospine mixture. Nano Lett., 2001, 1, 207–211.

21Yang, Ye; Rodriguez-Cordoba, W.; Xiang, X.; Lian, T. Strong Electronic Coupling and

Ultrafast Electron Transfer between PbS Quantum Dots and TiO2 nanocrystalline Films. Nano

Lett., 2012, 12, 303-309.

22Harel, E.; Rupich, S. M.; Schaller, R. D.; Talapin, D. V.; Engel, G. S. Measurement of

Electronic Splitting in PbS Quantum Dots by Two-dimensional Nonlinear Spectroscopy Phys.

Rev. B, 2012, 86, 075412.

57

23

Tang,J; Kemp,K. W.; Hoogland, S.; Jeong, K.S.; Liu, H.; Levina, L.; Furukawa, M.; Wang,

X.; Debnath, R.; Cha, D.; Chou, K.W.; Fischer, A.; Amassian, A.; Asbury, J.B.; Sargent, E. H.

Colloidal-quantum-dot photovoltaics using atomic-ligand passivation Nature Mater.2011, 10,

765–771

24 Jeong, K. S.; Tang, J.; Liu, H.; Kim, J.; Schaefer, A. W.; Kemp, K.; Levina, L.; Wang, X.;

Hoogland, S.; Debnath, R.; Brzozowski, L.; Sargent, E. H.; Asbury, J. B. Enhanced mobility-

lifetime products in PbS colloidal quantum dot photovoltaics ACS Nano, 2012, 1, 89-99

25 E. Talgorn;Gao, Y.; Aerts, M.; Kunneman, L. T.; Schins, J. M.; Savenije, T. J.; van Huis, M.

A., van der Zant, H. S. J.; houtepen, A. J.; Siebbeles, L. D. A. Unity Quantum Yield of

Photogenerated Charges and Band-like Transport in Quantum-Dot Solids. Nature Nanotech.2011,

6, 733.

26 Diaconescu, B.; Padilha, L. A.; Nagpal, P.; Swartzentruber, B. S.; Klimov, V. I. Phys. Rev.

Lett., 2013, 110, 127406

27

Zhang, J.; Jiang, X. Confinement-Dependent Below-Gap State in PbS Quantum Dot Films

Probed by Continuous-Wave Photoninduced Absorption. J. Phys. Chem. B, 2008, 112, 9557.

28Lewis, J. E.; Wu, S.; Jiang, J. Unconventional Gap State of Trapped Exciton in Lead Sulfide

Quantum Dots. Nanotechnology, 2010, 21, 455402.

29An, J. M.; Franceschetti, A.; Zunger, A.TheExcitonic Exchange Splitting and Radiative

Lifetime in PbSe Quantum Dots Nano Lett., 2007, 7, 2129-2135.

58

Chapter 4 Enhancement of mobility-lifetime product in PbS colloidal quantum dot solid.

This chapter is based upon the following publication.

Kwang S. Jeong, Jiang Tang, Huan Liu, Jihye Kim, Andrew W. Schaefer, Kyle Kemp, Larissa

Levina, Xihua Wang, Sjoerd Hoogland, RatanDebnath, Lukasz Brzozowski, Edward H. Sargent,

and John B. Asbury

Enhanced mobility-lifetime products in PbS colloidal quantum dot photovoltaics

ACS Nano, 2012, 1, 89-99

Kwang S. Jeong, Ryan D. Pensack and John B. Asbury

Vibrational Spectroscopy of Electronic Processes in Emerging Photovoltaic Materials

Accounts of Chemical Research, ASAP

This project was performed via collaboration with Dr. Edward Sargent's group in department of

computer and electrical engineering at the University of Toronto in Canada. The electrical

measurements and images of colloidal quantum dots were obtained by the Sargent group.

59

4.1 Introduction

The power conversion efficiency (PCE) of colloidal quantum dot (CQD) solar cells is

partially determined by: 1) the efficiency at which photons are absorbed and converted into

charge carriers, and 2) the quantum yield for collecting those charge carriers in an external

circuit. 12

These processes are optimized for different thicknesses of material and so device

developers should balance light absorption with charge collection. For instance, an active layer

thickness of 1 µm is necessary to efficiently absorb incident photons (90%) at the bandgap of

CQDs. However, charge harvesting efficiency increases as the active layer decreases in thickness

since charge carries are susceptible to recombination if their transit time exceeds the charge

recombination lifetime. An effective electrical length is therefore typically a few hundred

nanometers in current optimized CQD devices.

In order to improve photo harvesting efficiency at the bandgap energy, it is necessary to

increase the diffusion length of charge carriers. The diffusion length (l) is described as the below

equation,

𝑙 = 𝑘𝐵𝑇

𝑒µ𝜏

where kB, T, e, µ, τ, is Boltzmann's constant, temperature, elementary charge, the charge carrier

mobility, the charge carrier lifetime , respectively.

This chapter discusses how the ligand passivation strategy influences the mobility-

lifetime products of PbS CQD films and their relevance to device performance. We analyzed the

3-mercaptopropionic acid capped PbS CQD device (5.1 % PCE) and the ethanedithiol (EDT)

capped PbS CQD device (1.7%) using microsecond transient infrared spectroscopy (TRIR)

coupled with electrical measurements of minority carrier mobility (Figure 4-1).We found that the

superior power conversion efficiency of the MPA capped PbS CQD devices originates from

60

enhancement of the mobility-lifetime product which is a key parameter in diffusion length at

ambient condition. The discovery highlights the importance of ligands that efficiently passivate a

variety of surface states leading to strong interparticle coupling for efficient charge transport and

higher PCE in CQD solar cell devices.

4.2 Results and Discussion

When PbS CQDs are synthesized, they are capped with long oleic acid molecules for

protection and colloidal stability. Due to the insulating barrier of long alkyl chains surrounding

the nanoparticle, however, the charge carrier mobility of the oleic acid capped PbS CQD film is

significantly lower than what is needed for solar cell applications. To improve the mobility, the

long oleic acids are replaced with short ligands leading to increases in the carrier mobility.3,4,5,6,7,8

Figure 4-2 (a) represents the linear infrared spectra of oleic acid (OA), ethanedithoil

(EDT), 3-mercaptopropionic acid (MPA), propanedithiol (PrDT), butanedithiol (BDT),

pentanedithiol (PnDT), hexanedithiol (HDT), 1,2-benzenedithiol (1,2-BzDT) and 1,4-

benzenedithiol (1,4-BzDT) capped PbS CQD films, respectively. Among the various ligands

capped PbS CQD films, we focused on the MPA capped PbS CQD film and EDT capped PbS

CQD film which have 5.1 % and 1.7 % power conversion efficiency, respectively, in order to

analyze the correlation of the fundamental photophysical properties and the device performance.

In Figure 4-2 (b), the oleic acid capped PbS (dark blue) has two high intense peaks at

around 2900 cm-1

corresponding to the CH2 symmetric and asymmetric vibrational stretching

modes of the long alkyl chains. In addition, two other distinct peaks appear at 1405 cm-1

and

1545 cm-1

corresponding to the symmetric and asymmetric stretching modes of carboxylate

bound to the nanoparticle surface.9 The absence of a noticeable peak at 1720 cm

-1corresponding

61

to the unbound free carboxylic acid indicates that there is no residual free oleic acid and MPA in

the PbS CQD films.

By replacing the oleic acid with EDT, the intensities of the two peaks at around 2900 cm-

1 substantially reduce and the other two peaks at 1405 cm

-1and 1545cm

-1 reduce as well. The

intensity reduction of the vibrational peaks is indicative of the efficient ligand exchange to EDT

from the oleic acid. However, the MPA capped PbS CQD film still has the bound carboxylate

peaks at 1405 cm-1

and 1545 cm-1

while the intensities of CH2 peaks around 2900 cm-1

were

attenuated. This indicates that the bound oleic acid ligands were efficiently replaced by MPA.

All PbS CQD film thicknesses are approximately identical and the broad peak at 2350 cm-1

arises

from the carbon dioxide concentration in the spectrometer.

Scanning transmission electronic microscopy (STEM) was performed to examine how

ligand exchange reduces the interparticle distance.(Figure 4-3) Ligand exchange from oleic acid

to EDT or MPA reduces the interparticle distance, producing closely packed colloidal quantum

dot films. The MPA capped PbS film has an interparticle distance of 1.8 nm while the EDT film

has 2.1 nm, implying that electronic coupling in the MPA capped film will be more efficient than

that of EDT capped film. The interparticle distance was calculated by measuring average center-

to-center distance and subtracting the diameter of the PbS quantum dots. The quantum dots used

for the STEM were 5.0± 0.3 nm diameter PbS CQDs to efficiently visualize the interparticle

distance.

Figure 4-4 shows the microsecond transient infrared spectra (TRIR) of an EDT capped

PbS CQD film and that of a MPA capped PbS CQD film. The TRIR spectra reveal the temporal

transmission change of continuous wave infrared light before photoexcitation and at 0.5 µs after

photoexcitation by a 10 ns duaration 532 nm phoexcitation pulse. As discussed in Chapter 3, the

TRIR intensity reflects the density of charge carriers at the Stokes-shifted sub-gap state on

microsecond time scale.

62

There are two apparent features in the TRIR spectra: broad electronic transition at 0.25 –

0.35 eV; and narrow vibrational features of the ligands appearing at 0.15-0.2 eV (1200 – 1600

cm-1

). The broad electronic transition spreads over 0.15 - 0.50 eV and the transition energy is

around 0.3 eV. The most striking feature of the transient infrared spectra is the narrow vibrational

bleaches shown at 0.15 - 0.2 eV. The negative signed absorption feature at 1405 cm-1

and 1545

cm-1

represent the symmetric and asymmetric stretch of carboxylate functional group bound to

PbS quantum dot surface. The bleach is generated when the transmission is increased by

electrical field generated by charge carriers at the Stokes shifted sub-gap state which is coupled to

surface. As well as the bleaches, two other absorption peaks appear right next to the two bleach

features. We believe that the two absorption peaks arise from new absorption of electrically

perturbed carboxylate ligand vibrational stretching modes.

For better understanding the coupling between the sub-gap state and the surface of

quantum dot, we investigated the transition energy of the sub-gap to 1Pe state using various

ligands (Figure 4-5). The transition energy of the TRIR spectra was calculated by following

equation.

𝐸 = 𝐸𝑔 𝐸 𝑑𝐸

𝑔 𝐸 𝑑𝐸

where g(E) represents the broad absorption line shape in the TRIR spectra expressed in terms of

the transition energy, E.

Figure 4-5 shows the TRIR spectra of various ligand capped PbS CQD film: MPA, EDT,

PrDT, BDT, PnDT, HDT, 1,2-BzDT and 1,4-BzDT capped PbS CQD films. The red-shift of the

TRIR transition may be attributed to the reduction of Stokes shift by ligand variation.

Additionally, the near-IR excitonic absorption peaks were measured for the same samples (Figure

63

4-6). All near-IR absorption spectra are normalized at maximum absorption peak to compare the

bandgap energy. Interestingly, the transition energy of TRIR spectra of various ligands capped

PbS CQD films do not show strong correlation with the bandgap of PbS CQD films (Figure 4-7).

The weak correlation arises from the less surface sensitivity of bandgap transition while the TRIR

transition is sensitive to surface of CQD. The bandgap transition is mainly involved with the

core-character delocalized states whiles the TRIR transition is related to the surface coupled

Stokes shift sub-gap states.

The decays of the TRIR spectra of the MPA capped film and the EDT capped film were

measured at the maximum intensity (Figure 4-8). The averaged lifetimes of the EDT capped film

and the MPA capped film are 93 to 31 µs, respectively. The lifetime indicates the bimolecular

recombination lifetime of charge carrier at the sub-gap state.

Mobility measurements were performed by the Sargent group using ion gel field effect

transistor (FET).1011

Figure 4-9 shows the drain current of EDT capped PbS film and that of MPA

capped PbS film. Surprisingly, the drain current of the MPA sample is larger than that of the

EDT sample by an order of magnitude. The mobility of the EDT and MPA samples are 2.4 x 10-4

cm2/V·s and 5.1 x 10

-3cm

2/V·s, respectively.

In combination with the recombination lifetime, we found that the mobility-lifetime

product positively contributes to the device performance. It is worth to note that the mobility-

lifetime product is a key parameter determining the diffusion length of charge carriers at ambient

temperature. By changing the ligand from EDT to MPA, the mobility increased by a factor of 20

from 2.4 x 10-4

to 5.1x 10-3

. In contrast, the averaged lifetime measured by the transient infrared

spectroscopy only reduced by a factor of 3 from 93 to 31 µs. Therefore, the mobility-lifetime

product increased by a factor of 7 by varying the ligand from EDT to MPA. The enhancement of

the mobility-lifetime products was a surprising result considering conventional mean field theory.

According to mean field theory, the lifetime of a diffusion controlled bimolecular charge

64

recombination process is inversely proportional to the mobility.12

Therefore, if the charge

recombination trend in diffusion processes were following the mean field theory, the mobility-

lifetime product should not have been changed.

The enhancement of mobility-lifetime product was rationalized by comparing the

recombination lifetime of differently passivated films. Assuming that the trap states are evenly

spread on the surface of quantum dots, the space between trap states of the MPA capped film

should be larger than that of the EDT capped film because of the lower density of traps.

Therefore, due to the larger inter-trap space, the recombination process of the MPA capped film

is slower than what is estimated.

Based on the observation that the ligand replacement from EDT to MPA increases the

mobility-lifetime product, we examined the device performance of each film to find a correlation

with the mobility-lifetime product. In Figure 4-9, current density - bias voltage (I-V) curves were

measured under AM 1.5 solar illumination for each EDT and MPA capped PbS CQD film. The

ligand change from EDT to MPA gives rise to a dramatic improvement in the short-circuit

current, increasing from 7.4 to 17.5 mA/cm2 while the open-circuit voltage is increased by only

0.07 V. Consequently, the increase of both short-circuit current and open-circuit voltage leads to

significant improvement in the power conversion efficiency from 1.7 % to 5.1 %.

65

4.3 Conclusion

The combination of the microsecond transient infrared spectroscopy and electrical

measurements reveals why the device performance of PbS CQD film increases through surface

modification via ligand passivation strategy. The combined measurements show that the charge

recombination dynamics and mobility are substantially changed by varying the ligand passivation

strategy. The mobility- lifetime product, which is a key parameter of diffusion length, was

enhanced by a factor of 7 via ligand variation from EDT to MPA. The increase of mobility-

lifetime product may be attributed to the decrease of the density of charge traps. Along with the

enhancement of the mobility-lifetime product, the power conversion efficiency achieved 5.1 %

of MPA capped PbS CQD film.

66

Figure 4-1. Molecular structure of ligands: Oleic acid (OA), Ethanedithiol (EDT) and 3-

mercaptopropionic acid (MPA)

67

Figure 4-2. Infrared spectra of oleic acid (OA), 3-mercaptopropionic acid (MPA),

ethanedithoil (EDT), propanedithiol (PrDT), butanedithiol (BDT), pentanedithiol (PnDT),

hexanedithiol (HDT), 1,2-benzenedithiol (1,2-BzDT) and 1,4-benzenedithiol (1,4-BzDT) capped

PbS CQD films.

(A)

(B)

68

Figure 4-3. Scanning transmission electron microscopy (STEM) images of EDT capped

PbS CQD and MPA capped PbS CQD films. Average interparticle distance for the EDT capped

film and the MPA capped film are 2.1 nm and 1.8 nm, respectively.

(A)

(B)

69

Figure 4-4. Microsecond transient infrared (TRIR) spectra of EDT capped PbS film

(green) and MPA capped PbS film (blue). The narrow vibrational features result from ligands that

are electronically perturbed by charges carriers. The broad absorption features corresponds to

subgap-to-1Pe transition.

70

Figure 4-5. Microsecond transient infrared spectra of various ligand capped PbS films.

The microsecond transient infrared spectra is strongly dependent upon the ligand passivation. The

maximum transition energy shift is about 0.13 eV and the area of the spectra is correlated to the

length of ligand.

71

Figure 4-6.Excitonic absorption peaks of 3-mercaptopropionic acid (MPA), ethanedithoil

(EDT), propanedithiol (PrDT), butanedithiol (BDT), pentanedithiol (PnDT), hexanedithiol

(HDT), 1,2-benzenedithiol (1,2-BzDT) and 1,4-benzenedithiol (1,4-BzDT) capped PbS CQD

films. The bandgap energy is dependent on ligand passivation and the maximum shift is ~0.07

eV.

72

Figure 4-7.Mid-IR transition and bandgap correlation data. The scattered dots of various

ligand capped PbS CQD film implies the mid-IR transition is not strongly dependent on quantum

confinement.

73

Figure 4-8. TRIR decays of the EDT capped film (green) and the MPA capped film

(blue). The MPA capped film has lower lifetime by a factor of 3 in comparison to the EDT

capped film.

74

Figure 4-9. Id-Vg curves for minority carriers (electrons) in the EDT and MPA capped

PbS Films. The mobility was measured using the ion-gel field-effect transistor geometry. The

inset is the magnified Id-Vg curve for the EDT capped PbS film for clarity.

75

Table 4-1. Electrical and optical parameters of EDT capped PbS CQD film and MPA

capped PbS CQD film.

EDT MPA

µ<τ> product 0.022 0.16

µ (cm2/V s) 2.4 x 10

-4 5.1 x 10

-3

<τ> (µs) 93 31

< E > (eV) 0.29 0.27

𝐼 𝑣 𝑑𝑣 3.3 1.0

C-Vg (cm-3

) 8 x 1016

2 x 1016

76

4.5 Reference

1Tang, J.; Sargent, E. H. Infrared Colloidal Quantum Dots for Photovoltaics: Fundamentals and

Recent Progress. Adv. Mater. 2011, 23, 12–29.

2Talapin, D. V.; Lee, J.-S.; Kovalenko, M. V.; Shevchenko, E. V. Prospects of Colloidal

Nanocrystals for Electronic and Optoelectronic Applications. Chem. Rev. 2010, 110, 389–458.

3Tang, J.; Brzozowski, L.; Barkhouse, D. A. R.; Wang, X.; Debnath, R.;Wolowiec, R.; Palmiano,

E.; Levina, L.; Pattantyus-Abraham, A. G.; Jamakosmanovic, D.; Sargent, E. H. Quantum Dot

Photovoltaics in the Extreme Quantum Confinement Regime: The Surface-Chemical Origins of

Exceptional Air and Light-Stability. ACS Nano,2010, 4, 869–878.

4 Luther, J. M.; Law, M.; Beard, M. C.; Song, Q.; Reese, M. O.; Ellingson, R. J.; Nozik, A. J.

Schottky Solar Cells Based on Colloidal Nanocrystal Films. Nano Lett., 2008, 8, 3488–3492.

5 Luther, J. M.; Gao, J.; Lloyd, M. T.; Semonin, O. E.; Beard, M. C.; Nozik, A. J. Stability

Assessment on a 3% Bilayer PbS/ZnO Quantum Dot Heterojunction Solar Cell. Adv. Mater.,

2010, 22, 3704–3707.

6Choi, J. J.; Lim, Y.-F.; Santiago-Berrios, M. B.; Oh, M.; Hyun, B.-R.; Sun, L.; Bartnik, A. C.;

Goedhard, A.; Malliaras, G. G.; Abruna, H. D.; Wise, F. W.; Hanrath, T. PbSe Nanocrystal

Excitonic Solar Cells. Nano Lett., 2009, 9, 3749–3755.

7Leschkies, K. S.; Beatty, T. J.; Kang, M. S.; Norris, D. J.; Aydil, E. S. Solar Cells Based on

Junctions Between Colloidal PbSe Nanocrystals and Thin ZnO Films. ACS

Nano,2009, 3, 3638–3648.

8Kovalenko, M. V.; Scheele, M.; Talapin, D. V. Colloidal Nanocrystals with Molecular Metal

Chalcogenide Surface Ligands. Science, 2009, 324, 1417–1420.

77

9Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, X.; Konstantatos, G.;

Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, M. K.; Gratzel, M.; Sargent, E. H. Depleted-

Heterojunction Colloidal Quantum Dot Solar Cells. ACS Nano, 2010, 4, 3374-3380.

10Kang, M. S.; Lee, J.; Norris, D. J.; Frisbie, C. D. High Carrier Densities Achieved at Low

Voltages in Ambipolar PbSe Nanocrystal Thin-Film Transistors. Nano Letter,2009, 9, 3848-3852.

11Kang, M. S.; Sahu, A.; Norris, D. J.; Frisbie, C. D. Size-Dependent Electrical Transport in

CdSe Nanocrystal Thin Films. Nano Lett., 2010, 10, 3727-3732.

12Waite, T. R. Theoretical Treatment of the Kinetics of Diffusion-limited reactions. Phys.

Rev.1957, 107, 463-470.

78

Chapter 5 ATOMIC LIGAND PASSIVATED LEAD SULFIDE COLLOIDAL QUANTUM

DOT SOLID

This chapter was written on the basis of following publication.

1) Jiang Tang, KyleW. Kemp, Sjoerd Hoogland, Kwang S. Jeong, Huan Liu, Larissa

Levina,Melissa Furukawa, XihuaWang, Ratan Debnath, Dongkyu Cha, KangWei Chou, Armin

Fischer,AramAmassian, John B. Asbury and Edward H. Sargent Colloidal-quantum-dot

photovoltaics using atomic-ligand passivation

Nature Mater.2011, 10, 765–771

2) Kwang S. Jeong, Ryan D. Pensack and John B. Asbury Vibrational Spectroscopy of

Electronic Processes in Emerging Photovoltaic Materials

Acc. Chem. Res., ASAP

This project was conducted via collaborating with the Sargent group at the University of

Toronto in Canada. To note, the electrical measurements and the imaging quantum dots were

performed by the Sargent group.

79

5.1 Introduction

Ligand exchange in colloidal quantum dots (CQDs) from long insulating alkyl chains to

short size ligands has shown to minimize the interparticle distance to promote carrier transport

and lower the defect density to reduce recombination loss. The short alkylthiols12

, alkylamines3,

aromatic thiols4 and mercaptocarboxylic acid

5 treatment has shown promising achievement in

surface passivation, reduction of interparticle distant and device performance.

From the previous chapter, we learned that the ligand variation from EDT to MPA

enables charge carriers to undergo efficient transport by reduction of the density of trap states,

leading to high power conversion efficiency of 5.1 %. Nevertheless, alternative passivation

strategy is required due to the intrinsic limitation of the organic ligand: vulnerability to oxidation,

thermal degradation and insulating property. Through collaborative work with the Sargent group

at the University of Toronto in Canada, we first reported the atomic size inorganic ligand

passivated PbS colloidal quantum solar cell of 6% world record power conversion efficiency.

Furthermore, we investigated the fundamental photophysics determining the device

performance.6 The analysis of the atomic ligand passivated PbS colloidal quantum dot solid

provides better understanding of the electronic states of PbS quantum dot film and guiding future

direction for development strategies of CQD solar cell.

80

5.2 Result and discussion

Atomic ligand passivation was enabled using halide compound such as

cetyltrimethylammonium bromide (CTAB), hexadecyltrimethylammonium chloride (HTAC) and

tetrabutylammonium iodide (TBAI). The bulk halide solution was dispersed on the oleic acid

capped PbS CQD layer, therefore, the bulk ionic compound slowly releases the halide and

replaces the oleic acid bound to the quantum dot surface with the halide (Figure 5-1). Figure 5-2

represents the steady state infrared spectra of the oleic acid capped PbS CQD film and the halide

(Br-, Cl

-, I

-) capped PbS CQD film, displaying the change of the linear infrared transmission of

PbS CQD before and after ligand exchange. The asymmetric and symmetric stretching modes of

the CH2 alkyl chains ( 2922 cm-1

and 2852 cm-1

) and the carboxylate (1545 cm-1

and 1403 cm-1

)

disappear by ligand exchange to Br-, Cl

- and I

-, indicating that the bound oleic acids were

efficiently replaced by halides. The peaks appearing at 2345cm-1

arise from carbon dioxide

molecules in the spectrometer.

X-ray photoelectron spectroscopy (XPS) confirmed the existence of bromide of the

bromide capped PbS CQD film. (Table 5-1) The elemental atomic percentage of Pb 4f, S 2s, Br

3d were obtained and they are 48%, 33.8 %, and 17.8 %, respectively. The difference of the

elemental atomic percentage between Pb (4f) and S (2s) quantitatively matches the atomic

percentage of Br (3d) on the surface of the bromide capped PbS CQD.

Scanning transmission electron microscopy (STEM) images offer further evidence of the

ligand exchange. Figure 5-3 displays the STEM image of the oleic acid capped PbS CQD treated

by pure methanol (A), 1 % EDT in acetonitrile (B), 10 % MPA in methanol (C), and 10 mg/mL

CTAB in methanol (D). Center-to-center spacing of quantum dots was analyzed and the

interparticle distances were obtained: 2.1 nm for EDT capped film, 1.8 nm for MPA capped film

81

and 1.0 nm for bromide capped film. As observed in the image, ligand exchange reduces the

interparticle distance which is related to the size of the ligand molecule.

The ligand exchange also influences the near-IR absorption spectra. Figure 5-4 shows

that the first excitonic peak is red-shifted and broadened by ligand exchange. The peak

broadening and red-shift appear because quantum dots are more electronically coupled with

neighboring dots through more electron wavefunction overlapping. The red-shift and peak

broadening are also seen in other organic ligand capped PbS CQD films. Figure 5-5 represents the

near-IR absorption spectra of Br, MPA, 1,4-BzDT, BDT and EDT ligand capped PbS CQD film.

Interestingly, the longest ligand, 1,4-BzDT, produces the most narrow half-width-at-half

maximum(HWHM). The HWHM becomes broader by reducing the ligand length from

butanedithiol (BDT), ethanedithiol (EDT), 3-mercaptopropionic acid (MPA), to bromide (Br-). To

note, the bandgap absorption peak broadening and redshift do not result from the polydispersity

of size distribution since all samples were from the same CQD batch.

It has been reported that the minority carrier (electron) in the p-type PbS CQD film

determines the overall solar power conversion efficiency of the device. The electron mobility of

the p-type PbS CQD film was measured using ion-gel field effect transistor (Figure 5-6). The

atomic ligand passivation allowed to achieve 4 x 10-2

cm2/V·s in mobility which is an order of

magnitude larger than that of MPA capped PbS CQD film (5.1 x 10-3

cm2/V·s) and two orders of

magnitude larger than that of EDT capped PbS film (2.4x10-4

cm2/V·s).

Using the capacitance-voltage (C-Vg) measurement, a p-type doping (electron acceptor)

density of the bromide capped PbS CQD films was measured to be Nacpt= 3.7 x1015

cm-3

. The

doping density is one order of magnitude lower than that of MPA (2 x 1016

cm-3

) or EDT (8 x 1016

cm-3

) capped PbS CQD films, suggesting that the atomic passivation suppresses the doping states

including trap states,1 via an efficient atomic size ligand passivation strategy. The reduction of

density of states at around quasi-Fermi level (cm-3

/eV) implies the decrease of density of trap

82

states.7 Therefore, the efficient ligand passivation using halide reduces the intrinsic trap density

of the PbS CQD film, leading to higher mobility of electron.

The less density of trap states of the bromide capped PbS CQD film is probably attributed

to less exposure to oxygen in the ambient condition, resulting from the efficient atomic ligand

passivation. During the film preparation process, the ligand exchange is performed under

ambient condition, so the PbS CQD is exposed to air. Oxygen in air permeates into the PbS CQD

which was intrinsically n-type, converting the conductivity from n-type to p-type.

The enhanced mobility also can be partially explained by change of the dielectric

property. The dielectric constant (ε) of the bromide capped PbS CQD film was measured by

carrier extraction by linear increasing voltage (CELIV) technique and calculated by below

equation.(Figure 5-7)

Ɛ = 𝑗0𝑑

𝐴Ɛ0

where j0, d, A and Ɛ0 are the displacement current density, the film thickness, the ramp rate of

applied linearly increasing voltage and the vacuum permittivity, respectively (Figure 5-7).

The measured dielectric constant of bromide capped PbS CQD film is 25, which is larger

than 18 of ethanedithiol (EDT) capped film8 or 15 of 1,4-benzenedithiol (1,4-BzDT) capped

film.9 Interestingly, the reduction of the interparticle distance by ligand variation gives rise to the

increase of dielectric constant of PbS CQD. Since the dielectric constant of the bulk PbS film is

175,10

the gradual increase of the dielectric constant via reducing ligand size seems to be a

reasonable trend. The dielectric constant is related to the mobility by following equations.

Based on the capacitance (C) equation, the increase of dielectric constant (Ɛr) leads to the

enhancement of capacitance.

𝐶 = Ɛ0Ɛ𝑟𝐴

𝑑

83

where ,Ɛ0, Ɛr, A and d are the vacuum permittivity, the relative permittivity (dielectric

constant), the area of electrode and the thickness of the sample. Since the capacitance is inversely

correlated with the mobility through following equations, it is apparent that enhancement of

mobility originates from the increase of dielectric constant probably resulting from efficient

interparticle coupling.

𝑛 = 𝐶𝑉𝑔

𝐴

𝑉2

𝑉1

𝑑𝑉𝑔

and

µ = 𝐿𝐼𝑑

𝑊𝑒𝑉𝑑𝑛

where L, Id, W, e, n, A, C, V1 and V2 are the channel length, the drain current, the channel

width, the elementary charge, the accumulated carrier density in the channel, the area of the gate

electrode, the capacitance and the initial and final gate voltages, respectively.

Microsecond time scale transient infrared spectroscopy (TRIR) provides useful

information about the charge carrier recombination dynamics, the density of carriers on

microsecond time scale and the depth of sub-gap states in PbS CQD solid. The TRIR spectra of

the Br- capped film, the MPA capped film and EDT capped film are plotted in Figure 5-8. In the

spectra, two distinct features are observed: broad electronic transition; and vibrational features at

around 0.15-0.2 eV.

The maximum intensity of the TRIR spectra decreases in the order of EDT, MPA and Br

capped films. Given that the TRIR spectra reflect the density of charge carrier, the bromide

capped film has the smallest carrier density at 0.5µs after photoexcitation. The fast decay of the

bromide capped PbS CQD film is also shown in the inset of Figure 5-9.

It is interesting that the mid-IR transition energy is dependent upon the ligand and it is

correlated with the recombination lifetime and the mobility (Figure 5-9 and 5-10). The transition

84

energy difference originates from the different Stokes shift of the samples. The bromide capped

film shows the fastest recombination and the largest mobility. This is because the surface of the

bromide capped PbS CQD is efficiently passivated so the density of trap states is smaller than that

of MPA capped film or EDT capped film,7 leading to the fastest recombination and the largest

mobility.

The observation that the CQD with small density of traps shows small recombination

lifetime is against to the conventional Schokley-Read-Hall recombination theory that the material

with small density of traps shows large recombination lifetime. The opposite result is attributed

to the different types of traps. The Schokley-Read-Hall recombination assumes all traps to be

deep trap where charge carriers cannot escape from, however, the nanoparticle has mainly two

different trap states: shallow traps and deep traps also known as recombination centers.

Especially, the doping density is dominated by shallow traps where the trapped charge carriers

can be promoted to delocalized states when efficient external energy is applied. Therefore, there

is discrepancy between our result and the Schokley-Read-Hall recombination theory.

Other halide ligands were examined as well (Figure 5-11). Chloride (Cl-) capped PbS

CQD film, iodide (I-) capped PbS CQD film and the Cd

2+ treated bromide capped PbS CQD film

were measured and it turned out that they are not much different from the bromide.

In combination with the mobility measurement, mobility-lifetime products were obtained

for Br- capped PbS film, MPA capped PbS film and EDT capped PbS film. The mobility-lifetime

products of the ligand treated PbS CQD films are enhanced by a factor of 20 by varying the

ligands used to treat the films from EDT to bromide. Simultaneously, the power conversion

efficiency of the devices improved from 1.7 % (EDT) to 6.0% (Br-). Furthermore, the short-

circuit currents and the open circuit voltage were improved, primarily due to reduction of the trap

densities in the films.

85

The dramatic increase of the short-circuit current can be understood in terms of changes

of the exciton binding energy (Eb) and enhancement of diffusion length, which are involved with

charge carrier generation and collection. The exciton binding energy is involved with charge

generation. Since the exciton binding energy is inversely proportional to the dielectric constant of

material, an exciton of a material with large dielectric constant is efficiently dissociated to

produce charge carrier.

𝐸𝑏 =µ𝑒4

2ħ2Ɛ2=

ħ2

2µ𝑎𝐵2 and 𝑎𝐵 =

ħ2Ɛ

µ𝑒2

where µ, e, ħ, Ɛ, aB are the reduced mass, the elementary charge, the reduced Planck constant, the

dielectric constant (Ɛ0Ɛr) and the Bohr radius, respectively.10

The binding energies for the hybrid,

Br-, EDT and 1,4-BzDT capped PbS CQDs were obtained. The hybrid capped PbS CQD (Ɛ =

40), the Br capped PbS CQDs (Ɛ = 25), the EDT capped PbS CQDs (Ɛ =18) and the 1,4-BzDT

capped PbS CQDs (Ɛ =15) have the exciton binding energies of 1.34 meV, 3.42 meV, 6.60 meV

and 9.50 meV, respectively. By changing the ligand passivation strategy, the exciton binding

energy is varied due to the change of dielectric constant. Therefore, the increase of the short-

circuit current can be partially understood by the ligand dependent exciton binding energy.

Diffusion occurs in the quasi-neutral region where the built-in-potential is negligible.

This is the identical condition to the short circuit which has zero electrical potential. Therefore,

the increase of diffusion length enables more charge carriers to reach electrodes, leading to larger

short-circuit currents. Consequently, the increase of the short-circuit current results from the

decrease of the exciton binding energy related to carrier generation and from the enhanced

mobility-lifetime product involved with charge harvesting.

86

5.3 Conclusion

In conclusion, we revealed the fundamental photophysical properties of atomic ligand

passivated PbS CQD film, reasoning the world record high power conversion efficiency of 6.0%

in colloidal quantum dot solar cell.

Especially, the microsecond transient infrared spectroscopy offered information of not

only the charge recombination dynamics but also the density of charge carriers in the

microsecond time scale at the sub-gap state which is strongly coupled to the surface. By

combining the charge carrier mobility with the density of doping states, it was revealed that the

PbS CQD film with less trap density has faster recombination and higher mobility. The mobility-

lifetime products of the bromide capped film, the MPA capped film and the EDT capped film

were obtained and the mobility-lifetime product of the bromide shows a 20-fold increase

comparing to that of the EDT capped film. The enhancement of mobility-lifetime product was

reflected in the short-circuit current, the open-circuit voltage and the power conversion efficiency.

87

Figure 5-1. Atomic ligand passivation. Cd2+

cations of Cd-TDPA passivate S2-

dangling

bonds of PbS CQD at S1 step and the OA is replaced by bromide using CTAB at S2 step.

88

Figure 5-2. Infrared spectra of PbS CQDs before and after ligand exchange.

89

Figure 5-3. Scanning transmission electron microscopy (STEM) images of oleic acid

capped PbS CQD (A), EDT capped PbS CQD (B), MPA capped PbS CQD (C) and Br capped

PbS CQDs (D). From the images, the interparticle distances of each samples were measured: 2.1

nm of EDT capped PbS CQD; 1.8 nm MPA capped PbS CQD; 1.0 nm Br capped PbS CQD.

90

Figure 5-4. Near-infrared absorption spectra of oleic acid capped PbS CQD and

transmission spectra for Br, Cl and I capped PbS CQD film.

91

Figure 5-5. Near-infrared absorption spectra of Br, MPA, 1,4-BzDT, BDT and EDT

capped PbS CQD films. As decreasing ligand size, bandgap peak becomes broaden and red-

shifted.

92

Figure 5-6. Id-Vg (black) and C-Vg (red) curves of hybrid capped PbS CQD film.

93

Figure 5-7. Charge Extraction via a Linearly Increasing Voltage (CELIV) measurement

of Br capped PbS CQD film. Dielectric constant of the Br capped PbS CQD film was measured

to be 25±2.

94

Figure 5-8. Transient infrared spectra of Br, MPA and EDT capped PbS CQD films at

0.5 µs after photoexcitation with 532 nm pump pulse.

95

Figure 5-9. The average decay times of the trap-to-band transitions in PbS CQD films

are strongly correlated with the average mid-infrared transition energies

96

Figure 5-10.The mobilities of electrons in PbS CQD films are also strongly correlated

with the average mid-infrared transition energies.

97

Figure 5-11. Microsecond transient infrared spectra of halide capped PbS CQD film (Br,

Cl and I)

98

Table 5-1. X-ray photoelectron spectroscopy (XPS) analysis of elemental atomic

percentage

Sample Pb 4f S 2s Br 3d

Br- capped PbS 48.4 % 33.8 % 17.8 %

Table 5-2. Binding energy and dielectric constant correlation11

Ɛr Binding energy (meV) Approximate Radius (nm)

KCl 4.6 580 3

CuCl 5.6 190 7

Cu2O 7.1 150 7

Si 11.4 12 50

GaAs 13.1 4 150

99

5.4 Reference

1Klem, E. J. D.; Shukla, H.; Hinds, S.; MacNeil, D. D.; Levina, L.; Sargent, E. H. Impact of

Dithiol Treatment and Air Annealing on the Conductivity, Mobility and Hole Density in PbS

Colloidal Quantum Dot Solids. Appl. Phys. Lett., 2008, 92, 212105.

2Luther, J. M. et al Schottky Solar Cells Based on Colloidal Nanocrystal Films. Nano Lett., 2008,

8, 3488-3492.

3Talapin, D. V.; Murray, C. B. PbSeNanocrystal Solids for N- and P- Channel Thin Film Field-

Effect Transistors. Science, 2005, 310, 86-89

4Koleilat, G. I. et al Efficient, Stable Infrared Photovoltaics Based on Solution-cast Colloidal

Quantum Dots. ACS Nano, 2008, 2, 833-840.

5Pattantyus-Abraham, A. G.; Kramer, I. J.; Barkhouse, A. R.; Wang, X.; Konstantatos, G.;

Debnath, R.; Levina, L.; Raabe, I.; Nazeeruddin, M. K.; Gratzel, M.; Sargent, E. H. Depleted-

Heterojunction Colloidal Quantum Dot Solar Cells. ACS Nano, 2010, 4, 3374-3380.

6Tang,J; Kemp,K. W.; Hoogland, S.; Jeong, K.S.; Liu, H.; Levina, L.; Furukawa, M.; Wang, X.;

Debnath, R.; Cha, D.; Chou, K.W.; Fischer, A.; Amassian, A.; Asbury, J.B.; Sargent, E. H.

Colloidal-quantum-dot photovoltaics using atomic-ligand passivation. Nature Mater.2011, 10,

765–771

7 Ip, A. H.; Thon, S. M.; Hoogland, S.; Voznyy, O.; Zhitomirsky, R.; Debnath, R.; Levina, L.;

Rollny, L. R.; Carey, G. H.; Fischer, A.; Kemp, K. W.; Kramer, I. J.; Ning, Z.; Labelle, A. J.; Wei

Chou, K.; Amassian, A.; Sargent, E. H. Hybrid Passivated Colloidal Quantum Dot Solids Nature

Nanotech.,2012, 7, 577-582.

8Tang, J.; Brzozowski, L.; Barkhouse, D. A. R.; Wang, x.; Debnath, R.; Wolowiec, R.; Palmiano,

E.; Levina, L.; Pattantyus-Abraham, A.G.; Jamakosmanovic, D.; Sargent E. H. Quantum Dot

100

Photovoltaic in the Extreme Quantum Confinement Regime: The Surface-Chemical Origins of

Exceptional Air- and Light-Stability. ACS Nano, 2010, 4, 869-878.

9Koleilat, G. I.; Levina, L.; Shukla, H.; Myrskog, S. H.; Hinds, S.; Pattantyus-Abraham, A. G.;

Sargent, E. H. Efficient, Stable Infrared Photovoltaics Based on Sollution-Cast Colloidal

Quantum Dots ACS Nano, 2008, 2, 833-840.

10 Dvorak, M.; Wei, S. H.; Wu, Z. Origin of the Variation of Exciton Binding Energy in

Semiconductors. Phys. Rev. Lett., 2013, 110, 016402.

11Snoke, D. W. Solid State Physics: Essential Concepts; Addison-Wesley: San Francisco, 2009.

101

Chapter 6 Improvement of charge transport in organic photovoltaic

materials by tuning dielectric properties

6.1 Introduction

We learned from previous chapters that the dielectric property of colloidal quantum dots

used as photovoltaic material influences the charge carrier mobility. In this work, we investigated

charge carrier dynamics by varying the dielectric property in organic photovoltaic materials.

Organic photovoltaic materials consist of an electron donor and an electron acceptor material.

Charge carriers are generated in the electron donor domain under photoexcitation and transfer

into electron acceptor domain. In the transport process, not all charge carriers move into the

electron acceptor domain. Some carriers recombine through geminate and bimolecular

recombination processes, dissipating the photon energy either radiatively or non-radiatively. The

recombination process is a significant loss mechanism in solar cell device performance.

Recombination processes can be reduced by limiting the thickness of organic photovoltaic active

layers to approximately 100 nm.1,2,3,4

The transit time of charge carriers is faster than the

recombination lifetime in such a thin layer, enabling efficient charge collection.5 However, such

a thin layer is not desirable for charge generation because a thicker layer is required to absorb all

of the incident photons. In particular, the extinction coefficient of the electron donor significantly

decreases at long wavelength range. The regioregularpoly(3-hexylthiophene) (P3HT) and

phenyl-C61-butyric acid methyl ester (PCBM) organic photovoltaic material are exceptions since

thicker active layers are available due to the relatively long recombination lifetime despite of the

high carrier mobility. This behavior, known as non-Langevin behavior, has been sought to

improve device performance of organic photovoltaic materials. According to the Langevin

theory,6 bimolecular recombination is determined by an effective lifetime (τ)

102

1/

/ ( (0) )n

where γ is an empirical disorder parameter, n(0) is the initial charge carrier density and β

is the Langevin coefficient. The Langevin coefficient β is proportional to the spatially averaged

carrier mobility divided by the spatially averaged dielectric permittivity, / . The

lifetime is dependent upon an empirical disorder parameter, γ, and inversely correlated to the

initial charge carrier density, n(0). When assuming that the disorder parameter is not significant

(γ =1), the mobility-lifetime product (µτ) is proportional to the spatially averaged dielectric

constant (ε) and inversely correlated to the initial carrier density (n(0)). According to the

equations of diffusion ( 𝑙𝑑𝑖𝑓𝑓𝑢𝑠𝑖𝑜𝑛 = 𝑘𝑇µτ

𝑒 ), the mobility-lifetime product (µτ) determines the

diffusion length at room temperature. Hence, an increase in mobility-lifetime product enables a

thicker active layer in the OPV device, leading to higher power conversion efficiency (PCE).

Here, we investigated bimolecular charge recombination processes using microsecond

transient infrared spectroscopy (TRIR).7 In order to analyze how the mobility-lifetime product

influences the OPV device performance, charge carrier mobilities were measured via

collaborative work with Dr. Enrique Gomez's group in the Department of Chemical Engineering

at Penn State University. Furthermore, the dielectric properties of the polymer blends were finely

tuned for analysis of the dielectric change effect on the mobilty-lifetime product and device

performance. Fine dielectric property control was possible using a specially synthesized polymer

named M-TQ1 (Figure 6-1) synthesized by collaborators in Dr. Qing Wang’s group in the

Department of Materials Science and Engineering at Penn State University. The M-TQ1 has

oligo-ethylene oxide chains to allow the Li ion to spread uniformly in it. Strikingly, we observed

a 35-fold enhancement of the mobility-lifetime product by increasing the dielectric constant by

103

addition of the Li ion salt. The increase in both recombination lifetime and mobility is consistent

with studies of Grätzel and coworkers with solid state dye sensitized solar cells.8910

6.2 Results and Discussion

M-TQ1 is the semiconducting polymer used as the electron donor in this work. The

backbone structure of M-TQ1 was inspired by the work of Andersson and coworkers who

reported 6% power conversion efficiency using an octyloxy side chain polymer.11

The

polyethylene glycol side chains of M-TQ1 are tethered to the conjugated backbone structure for

uniform distribution of bis(trifluoro-methyl-sulfonyl)imide lithium salt (LiTFSI) ionic compound.

Dielectric functions for LiTFSI doped M-TQ1 film with varying doping density were measured

by using impedance spectroscopy at frequency range of 2-20 MHz (Figure 6-2). The relative real

dielectric function was calculated from the below equation,

2 2' '/ ( ") ( ')z C z z

where z' is the real impedance, is the angular frequency, C is the capacitance of the empty cell,

(0 /C A d , o is the permittivity of vacuum, A is the area of the electrode, d is the thickness of

the sample) and z'' is the imaginary impedance. Figure 6-2 (A) displays the real dielectric

functions of M-TQ1 polymer films with Li+/O ratio of 0, 0.01 and 0.015. The static dielectric

constant of pure M-TQ1 is about 4 at 1 kHz, which is slightly greater than that of P3HT.12

The

increase of the dielectric constant is caused by the oligo-ethylene oxide side chain of M-TQ1.

Addition of the Li ion to the M-TQ1 polymer enhances the dielectric constant to ~140.

Interestingly, while the static dielectric constant does not noticeably change at 1 kHz with

increase of Li ion concentration from Li+/O of 0.010 to Li

+/O of 0.015, the dielectric function of

the highest Li+/O (0.015) is larger than that of the lowest Li

+/O (0.010) at high frequency range. It

104

is worth to note that the sample films were kept under nitrogen atmosphere to avoid oxygen or

water exposure so the dielectric feature of water molecule, the rapid increase at lower frequency

due to electrode polarization, is not seen at the measured dielectric functions.13

Figure 6-2(B) represents the real impedance functions of the pure and the Li ion doped

(0.010 and 0.015 in Li+/O) M-TQ1 films. The pure M-TQ1 film shows much higher impedance

than that of doped polymers at low frequency (< 1kHz), meaning that the charge carrier density is

relatively lower than that of Li ion doped films. With increase of Li ion concentration from 0 to

0.010, the real impedance is reduced, reflecting the increase of carrier concentration in the film.

With all observations, it was confirmed that the LiTFSI addition to M-TQ1 increases the

dielectric permittivity of the M-TQ1 polymer film.

Mobilities of the pure and the Li ion doped M-TQ1 films were measured using a thin film

transistor. Figure 6-3 shows the source-drain current ISDas a function of gate voltage Vgfor 0,

0.05, 0.010 and 0.015 Li+/O M-TQ1 films. ISD increases in both negative and positive gate

voltage range as Li ion concentration increases. The calculated mobility at VSD = -50V indicates

that the hole mobility increases by two orders of magnitude at 0.015 Li+/O.

Bimolecular recombination kinetics in the M-TQ1 PCBM polymer blend were examined

using TRIR spectroscopy (Figure 6-4).The compound ratio of M-TQ1and PCBM is 1:3 which is

the optimized ratio for the best PCE (1.1 %). In TRIR, a 10 ns duration 532 nm pump excites the

polymer blend film and broad featureless polaron absorption in the mid-IR region is probed at 0.5

µs after photoexcitation(inset of Figure 6-4).14

The featureless polaron kinetic decay was

measured at 1700 cm-1

. The lifetime increases with increasing salt concentration: 16 µs for 0

(Li+/O),101µs for 0.005 (Li

+/O),119µs for 0.010 (Li

+/O), and 133µs for 0.015 (Li

+/O). Multi-

exponential fit functions were used to obtain the kinetic data and the fit functions were

convoluted with the instrument response function. The mobility and bimolecular recombination

lifetime are plotted in Figure 6-5 (A). By varying the Li ion concentration, the mobility changes

105

by over two orders of magnitude and the bimolecular recombination lifetime varies by a factor of

8. Mobility-lifetime products are obtained for each sample and plotted in Figure 6-5

(B).Surprisingly, the mobility-lifetime product shows proportionality to the Li+/O concentration,

indicating that the increase of dielectric constant leads to larger mobility-lifetime product.

Steady state visible absorption and photoluminescence (PL) emission spectra were

obtained for various Li ion concentration films. Figure 6-6 displays the optical bandgap

absorption spectra of pure M-TQ1 film, 0.010 Li/O doped M-TQ1 film, M-TQ1 PCBM 1:1 blend

and 0.010 Li/O doped M-TQ1 PCBM blend film. The result suggests that neither the Li ion salt

addition nor PCBM addition makes any difference in the bandgap absorption spectra. The

insensitive bandgap absorption implies that the conjugated framework of the polymer is not

significantly perturbed by formation of Li+-poly ethylene glycol complex. In contrast, the PL

emission spectra in Figure 6-7 are very sensitive to the Li ion addition and, moreover, the PCBM

addition induces blue-shift of 8 nm. The blue-shift is frequently observed in conjugated polymer

blends and has been attributed to disruption of inter-chain interaction in the polymer. The

discrepancy between the absorption spectra and PL spectra can be explained by sensitivity for

irregular structure. The absorption spectra are not sensitive to the surface defect or morphological

disruption since the excitonic transition is mostly involved with delocalized states of the

conjugated backbone. The PL emission, however, is capable of reflecting the irregular structure

effect since it is sensitive to the localization state arising from branch structure. Therefore, we

believe that the difference in bandgap absorption and emission spectra results from the polaron

generated by inhomogeneous disruption of polymer blend morphology due to the addition of Li

salt. We believe that the blue-shift in PL may result from the shift of the effective Fermi energy

closer to the valence band due to the increase of density of positive polaron.

Time-correlated single photon counting (TCSPC) experiments were conducted to

examine the effect of dielectric property change in exciton lifetime in the M-TQ1 film. The

106

exciton lifetime of pure M-TQ1 film and 0.010 Li+/O doped M-TQ1 film were measured to be 82

and 77 ps, respectively. Considering the doping density of the Li ion (1020

cm-3

), the small change

of exciton lifetime is surprising comparing with prior results with charged species in organic

semiconductors.1516171819

To note, prior studies examined ionic species that arose form or gave rise

to defects in the molecular structures of the organic materials. For instance, Gregg and co-

workers discovered a significant density of p-type and n-type defects in P3HT layers originating

from polymer conformations that strain the π-bond system of the conjugated backbone. The p-

type defects decreased by 400-fold, leading to a doubling of the exciton diffusion length and an

order of magnitude increase in the hole mobility. Similarly, charge defects due to injection of

holes or migration of metal ions in AlQ3 layers in OLEDs give rise to disruption of the molecular

structure of the light emitting layers with corresponding loss of electro-and PL efficiency. As

opposed to works by others, our Li ion doping does not produce significant changes in the

conjugated framework of the polymer as exhibited by the unchanged absorption spectra after

adding Li ion salt. The small exciton lifetime change is consistent with negligible perturbation of

the conjugated framework observed in the absorption spectra.

The morphology of the M-TQ1 PCBM blend was examined by using energy filtered

transmission electron microscopy (TEM). Figure 6-8displays a carbon map of the M-TQ1 PCBM

1:3 polymer blend. The light and dark gray domains represent carbon-rich domains corresponding

to the PCBM domain and the M-TQ1 polymer domain, respectively. Interestingly, the M-TQ1

polymer domain is significantly separated from the PCBM domain in macroscale, which is not

desired. The approximately 500 nm scale phase separation implies that only about 10 % of the M-

TQ1 in the blends is within the ~10 nm exciton diffusion length of the PCBM-rich phase.

Therefore, the addition of PCBM has not a significant impact on approximately 90 % of the M-

TQ1 in the polymer blend. The PL decay of the blend film appearing in Figure 6-7 should be

107

dominated by the majority component of the polymer that is center of the macroscale PCBM

domain.

In order for efficient charge separation at the interface, the M-TQ1 and PCBM should be

homogeneously distributed. The lack of interface area between M-TQ1 and PCBM is reflected in

the device performance. The M-TQ1 PCBM 1:3 polymer blend film achieved 1.1% power

conversion efficiency (PCE). That is, the addition of Li ion salt did not result in an increase of

PCE. The inconsistency between the mobility-lifetime product and the PCE is attributed to the

lack of interface area between M-TQ1 and PCBM domains.

108

6.3 Conclusion

In conclusion, the dielectric property of M-TQ1 was investigated using various optical

and electrical methods. The dielectric property of M-TQ1 polymer film was varied by addition of

lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) serving as the lithium ion source to the

electron donor polymer. M-TQ1 is optimally designed for the uniform distribution of the lithium

ion by tethering a 10 ethylene glycol structure to the backbone structure.

The dielectric property was quantified by using impedance spectroscopy. In order to

scrutinize the influence of dielectric property change on the optical property of M-TQ1, time-

correlated single photon counting, microsecond transient infrared spectroscopy, steady state

visible absorption and photoluminescence spectroscopy were utilized.

In combination with thin-film-transistor technique, the mobility-lifetime products of pure

M-TQ1 film and Li ion doped M-TQ1 films were measured. The mobility-lifetime product was

substantially enhanced with increasing dielectric constant. The energy filtered transmission

electron microscopy (TEM) demonstratedmacroscale phase separation in the M-TQ1 and PCBM

polymer blend, which is the reason why the device performance is not correlated with the

mobility-lifetime product. This result suggests that the mobility-lifetime product is very

sensitivetothemorphology of the material. For future work, it will be necessary to enhance the

interaction between the hydrophilic side chain and the hydrophobic PCBM molecule by addition

of more polarizable amphiphilic ionic compound or by attenuating the hydrophilicity of side

chain to minimize macroscale phase separation.

109

Figure 6-1.Structure of M-TQ1, Lithium Bis(Trifluoromethanesulfonyl)Imide (LiTFSI)

and Phenyl-C61-Butyric-Acid-Methyl Ester (PCBM).

M-TQ1 LiTFSI PCBM

110

Figure 6-2. Real dielectric functions (A) and real impedance functions (B) of M-TQ1

(red), 0.010 Li+/O (black), 0.015 Li+/O (green) doped M-TQ1 films.

111

Figure 6-3. Source-drain current, ISD, versus gate voltage, Vg, measured in thin films of

M-TQ1 doped with various concentrations of LiTFSI. The off current at positive gate voltage

increases substantially with lithium ion doping.

112

Figure 6-4. Charge recombination kinetics decays of 1:1 by mass of M-TQ1 PCBM

polymer blend doped with various concentrations of lithium ions. The transient signal arises

from the polaron absorption in the polymer phase measured at 1700 cm-1

, shaded region within

the inset. The gray curve represents the instrument response function (IRF) used to deconvolute

the kinetic decays. The data indicate slower charge recombination dynamics with increasing

lithium ion concentration.

113

Figure 6-5.(A)Li ion dependent mobility and bimolecular recombination lifetime of M-

TQ1 polymer film.(B)Both the hole mobility and recombination lifetime give rise to large

increases of the mobility-lifetime product.

114

Figure 6-6.The absorption spectrum of a film of Li-doped M-TQ1 is almost identical to

that of pure M-TQ1.

115

Figure 6-7. Steady-state fluorescence spectra of the pure M-TQ1, 0.010 Li+/O doped M-

TQ1, M-TQ1/PCBM 1:1 blend by mass and 0.010 Li+/O doped M-TQ1/PCBM 1:1 blend by

mass.

116

Figure 6-8. Time correlated single photon counting decays of M-TQ1 films in the

presence and absence of lithium ions are displayed. Interestingly, the exciton lifetime in M-TQ1

is not significantly affected by the presence of Li ions. Addition of PCBM to the pure polymer

film causes a modest decrease in the exciton lifetime.

117

Figure 6-9. Energy filtered TEM image of a 1:3by massM-TQ1 PCBM polymer blend.

The image shows a carbon map such that the lighter gray regions correspond to the PCBM-rich

phase and the dark gray correspond to the M-TQ1 domain. The image represents that M-TQ1

and PCBM undergo macroscopic phase separation.

118

6.4 Reference

1Chen, M.-H.; Hou, J.; Hong, Z.; Yang, G.; Sista, S.; Chen, L.-M.; Yang, Y. Efficient Polymer

Solar Cells with Thin Active Layers Based on Alternating Polyfluorene Copolymer/Fullerene

Bulk Heterojunctions. Adv. Mater. 2009, 21, 4238-4242.

2Chen, H.-Y.; Hou, J.; Zhang, S.; Liang, Y.; Yang, G.; Yang, Y.; Yu, L.; Wu, Y.; Li, G. Polymer

Solar Cells with Enhanced Open-circuit Voltage and Efficiency. Nature Photonics 2009, 3, 649-

653.

3Zhang, Y.; Hau, S. K.; Yip, Y.-L.; Sun, Y.; Acton, O.; Jen, A. K.-Y. Efficient Polymer Solar

Cells Based on the Copolymers of Benzodithiophene and Thienopyrroledione. Chem.

Mater.2010, 22, 2696-2698.

4Wienk, M. M.; Turbiez, M.; Gilot, J.; Janssen, R. A. J. Narrow-Bandgap Diketo-Pyrrolo-Pyrrole

Polymer Solar Cells: The Effect of Processing ont he Performance. Adv. Mater.2008, 20, 2556-

2560.

5Park, S.; Roy, A.; Beaupre, S.; Cho, S.; Coates, N.; Moon, J. S.; Moses, D.; Leclerc, M.; Lee, K.;

Heeger, A. J. Bulk Heterojunction Solar Cells with Internal Quantum Efficiency Approaching

100% Nature Photonics 2009, 3, 297-303.

6Langevin, P. Ann. Chem. Phys.1903, 28, 289-384.

7Pensack, R. D.; Banyas, K. M.; Asbury, J. B.; Charge Trapping in Organic Photovoltaic

Materials Examined with Time Resolved Vibrational Spectroscopy J. Phys. Chem.C, 2010, 114,

5344-5350.

8Kruger, J.; Plass, R.; Cevey, L.; Piccirelli, M.; Gratzel, M. High Efficiency Solid-state

Photovoltaic Device due to Inhibition of Interface Charge Recombination. Appl. Phys. Lett.2001,

79, 2085-2087.

119

9Snaith, H. J.; Gratzel, M. Enhanced Charge Mobility in a Molecular Hole Transpoter via

Addition of Redox Inactive Ionic Dopant: Implication to Dye-sensitized Solar Cells. Appl. Phys.

Lett.2006, 89, 262114(262113).

10Snaith, H. J.; Schmidt-Mende, L.; Gratzel, M. Light Intensity, Temperature, and Thickness

Dependence of the Open-circuit Voltage in Solid-state Dye-sensitized Solar Cells. Phys. Rev. B

2006, 74, 045306 (045306).

11Wang, E.; Hou, L.; Wang, Z.; Hellstrom, S.; Zhang, F.; Inganas, O.; Andersson, M. R. An

Easily Synthesized Blue Polymer for High-Performance Polymer Solar Cells. Adv. Mater.2010,

22, 5240-5244.

12Knipper, M.; Parisi, J.; Coakley, K. M.; Waldauf, C.; Brabec, C. J.; Dyakonov, V. Zeitschrift

Naturforsh.2007, 62, 490-494.

13Klein, R. J.; Zhang, S.; Dou, S.; Jones, B. H.; Colby, R. H.; Runt, J. Modeling Electrode

Polarization in Dielectric Spectroscopy: Ion Mobility and Mobile Ion Concentration of Single-ion

Polymer Electrolytes. J. Chem. Phys.2006, 124, 144903-144908.

14Sheng, C.-X.; Tong, M.; Singh, S.; Vardeny, Z. V. Experimental Determination of the

Charge/neutral Branching Ratio ƞ in the Photoexcitation of π-conjugated Polymers by Broadband

Ultrafast Spectroscopy. Phys. Rev. B 2007, 75, 085206(085207).

15Liang, Z.; Nardes, A.; Wang, D.; Berry, J. J.; Gregg, B. A. Defect Engineering in π-Conjugated

Polymers. Chem. Mater.2009, 21, 4914-4919.

16Aziz, H.; Popovic, Z. D. Degradation Phenomena in Small-Molecule Organic Light-Emitting

Devices. Chem. Mater.2004, 16, 4522-4532.

17Kondakov, D. Y.; Sandifer, J. R.; Tang, C.; Young, R. H. Nonradiative Recombination Centers

and Electrical Aging of Organic Light-emitting Diodes: Direct Connection Between

Accumulation of Trapped Charge and Luminance Loss. J. Appl.Phys. 2003, 93, 1108-1119.

120

18Popovic, Z. D.; Aziz, H.; Ioannidis, A.; Hu, N.-X.; dos Anjos, P. N. M.Time-resolved

Fluorescence Studies of Degradation in tris(8-hydroxyquinoline) Aluminum (AlQ3)-based

Organic Light Emitting Devices (OLEDs). Synthetic Metals 2001, 123, 179-181.

19Lee, S. T.; Gao, Z. Q.; Hung, L. S. Metal Diffusion from Electrodes in Organic Light-Emitting

Diodes. Appl. Phys. Lett. 1999, 75, 1404-1406.

121

Chapter 7 Conclusion

7.1 Summary

In conclusion, we found how the fundamental photophysical properties of photovoltaic

materials influence solar cell device performance using optical and electrical methods. In

chapters3,4 and 5, various optical methods including UFIR and TRIR spectroscopy were

discussed that reveal the origin of Stokes-shifted sub-gap state in PbS CQD solid. Investigations

combining with temperature dependent TRIR, transient photoconductivity and short-circuit

current measurements revealed that the Stokes-shifted sub-gap state is the charge transport state

in the highest efficiency PbS CQD solar cells.

The mobility-lifetime products were measured for various ligand capped PbS CQD

solids. It was found that efficient ligand passivation strategies lead to increased mobility-lifetime

products and device performance. The mobility-lifetime product determines the diffusion length

in the quasi-neutral region, which in turn affects device performance. By comparing the density

of trap states, it was realized that the mobility-lifetime product is inversely proportional to the

density of trap states.

In chapter 6, we examined the dielectric permittivity dependent mobility-lifetime product

of organic photovoltaic materials.M-TQ1, a polymer specially designed for this work, was used

as the electron donor polymer and PCBM was used as the electron acceptor molecule. We found

that the increase of dielectric permittivity of M-TQ1 PCBM polymer blend by addition of LiTFSI

salt gives rise to dramatic enhancement in mobility-lifetime product. However, the enhancement

of mobility-lifetime product was not reflected in the power conversion efficiency. The macroscale

phase separation in the M-TQ1 PCBM polymer blend arising from the 10 oligo-ethylene oxide

122

was the main reason for the discrepancy between the mobility-lifetime product and the device

performance.

7.2 Future direction

7.2.1 Bias controlled transient two dimensional infrared spectroscopy (BT2DIR)

As shown in chapters 3, 4 and 5, the optical and electrical properties are very sensitive to

the ligand passivation strategy. Therefore, it is critical to scrutinize the surface environment of

colloidal quantum dots using especially high sensitive optical approach such as the two

dimensional infrared (2D-IR) spectroscopy. Dimercaptoamide (DMA), which is not

commercially available and so was synthesized by the author, includes an amide bond in the

middle of the molecule and two thiols at each end. The amide bond is frequently used in biology

since it is the base of protein structure. Reports on protein folding dynamics using 2D-IR

spectroscopy show that vibrational modes of the amide bond are very sensitive to the surrounding

electrical potential, and therefore, the amide bond is frequently used as a probe.1Similarly, we

could examine vibrational modes of DMA to probe the surface environment of colloidal quantum

dot solids in the presence and absence of charge carrier, which will enable us to better understand

the carrier transport and trapping mechanism.

Preliminary result

The DMA ligand was synthesized through the following synthesis scheme.

123

124

Figure 7-1 represents the H-NMR of DMA molecule. The triplet, quartet, broad peaks of

DMA are present in the spectrum. The linear infrared absorption spectrum of the DMA capped

PbS CQD film was obtained as presented in Figure 7-2. The peaks at 1650cm-1

and 1550cm-1

represent C=O stretching mode of DMA (Amide I) and N-H bending mode (Amide II) of DMA

molecule. The H-NMR and linear IR spectra indicate that the DMA molecule was successfully

synthesized and bound to the surface of PbS CQDs via ligand exchange process.

The broadband infrared pump and infrared probe spectroscopy was performed to examine

the DMA capped PbS CQD films. A 200 fs duration mid-IR pump pulse excites the vibrational

manifold of the amide bond in the DMA molecule followed by a 200 fs duration mid-IR probe

pulse interacting with the excited DMA molecule. The bleach feature of the amide bond appears

at 1658 cm-1

and the new absorption peak arising from the perturbed DMA ligand is observed at

1620 cm-1

. The 1658 cm-1

indicates that the amide I bond is randomly bound to the surface of PbS

CQDs.

The broadband infrared pump and probe spectrum shows the possibility of using the

DMA molecule as a probe for bias controlled transient two dimensional infrared spectroscopy

(BT2DIR) in Figure 7-3. The BT2DIR is probing the charge transfer and trapping dynamics using

the electrical potential sensitive DMA probing molecule. By varying the Fermi energy level, the

bleach of DMA is expected to change under photoexcitation. Since it was revealed that the Stokes

shifted sub-gap state is coupled to the surface, the coherence between the Stokes-shifted sub-gap

state and the trap state (surface) should be probed.

125

The device geometry and setup for the BT2DIR are illustrated in Figure 7-4. The sample

cell consists of the semi-infrared transparent palladium electrode/ PbS CQD film/ Au electrode.

7.2.2 Time-resolved photoluminescence spectroscopy (TRPL)

TRPL will be able to probe the electron and hole recombination process through femto

to nano second time scale where a variety of photophysical events occur. The TRPL measurement

will clarify how the hole contributes to the photocarrier dynamics at very fast time scale

comparing to the result of ultrafast transient infrared spectra probing mainly electron dynamics.

TRPL is also capable of providinghow the surface passivation is involved in hole dynamics. In

combination with ultrafast transient infrared spectroscopy, it will be able to quantify how the

electron and hole dynamics are affected by ligand passivation.This approach was inspired by the

work done by Victor KilimovonCdSe/ZnS colloidal quantum dots. 2

126

Figure 7-1. H-NMR of dimercaptoamid (DMA)

127

Figure 7-2. Linear infrared spectrum of DMA

128

Figure 7-3. Broad infrared pump probe spectrum of DMA capped PbS CQD film.

129

Figure 7-4. Device geometry of BT2DIR

130

7.3 Reference

1DeFlores, L. P.; Ganim, Z.; Nicodemus, R. A.; Tokmakoff, A. Amide I'-II' 2D IR Spectroscopy

Provides Enhanced Protein Secondary Structural Sensitivity J. Am. Chem. Soc., 2009, 131, 3385-

3391.

2Klimov, V. I.; Mikhailovsky, A. A.; McBranch, D. W. Mechanisms for Intraband energy

relaxation in semiconductor quantum dots: The role of electron-hole interactions Phys. Rev.

B,2000, 61, 13349-13352)

131

Appendix

Extension power law fit function

𝑓 𝑡 = 𝐴1 × 1 − exp −𝑡

𝜏1 − 𝐴2 × 1 − exp −

𝑡

𝜏2 + (1 − 𝐴1 − 𝐴2)

× 1

𝑡

𝜏3 𝛾

+ 1

where A1, A2 are prefactor of two rising fit functions. Since the photophysical

phenomenon is associated with two charge carriers, power law function was used for fitting.γ is

the extention power representing the inhomogeneity of quantum dot.

VITA

Kwang Seob Jeong

Education

B.S. in Chemistry, Korea University, Seoul, Korea (2007)

2008-2013 Chemistry Ph.D., The Pennsylvania State University, University Park, PA,

USA Advisor: Prof. John B. Asbury

Publication

Kyle W. Kempǂ, Kwang S. Jeong

ǂ, Jihye Kim, OleksandrVoznyy, Sjoerd Hoogland,

Susanna M. Thon, Alex H. Ip, Robert J. Stewart, Edward H. Sargent and John B. Asbury

Stokes Shifted Sub-gap State Mediate Photocarrier Transport in Colloidal Quantum Dot

Solids. Submitted to Nature Communications.

Kwang S. Jeong, Jiang Tang, Huan Liu,, Jihye Kim, Andrew W. Schaefer, Kyle Kemp,

Larissa Levina, Xihua Wang, Sjoerd Hoogland, RatanDebnath, Lukasz Brzozowski,

Edward H. Sargent, and John B. Asbury Enhanced Mobility-Lifetime Products in PbS

Colloidal Quantum Dot Photovoltaics.

ACS Nano, 2012, 1, 89-99

Kwang S. Jeong, Ryan D. Pensack, John B. Asbury Vibrational Spectroscopy of

Electronic Processes in Emerging Photovoltaic Materials.

Account Chemical Research ASAP

Jiang Tang, Kyle M. Kemp, Sjoerd Hoogland, Kwang S. Jeong, Huan Liu, Larissa

Levina, Melissa Furukawa, Xihua Wang, Ratan Debnath, Dongkyu Cha, Kang Wei

Chou, Armin Fischer, Aram Amassian, John B. Asbury, Edward H. Sargent

Colloidal Quantum Dot Photovoltaics Using Atomic Ligand Passivation.

Nature Materials 2011, 10, 765-771