166
SOLUTION PROCESSED ORGANIC SEMICONDUCTOR THIN-FILM TRANSISTORS FOR FLEXIBLE ELECTRONICS: DEVICE PHYSICS, DEVICE MODELING, FABRICATION TECHNOLOGY, AND INTERFACE ENGINEERING A DISSERTATION SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Zihong (Bill) Liu November 2009 [Defense Date: September 10 th , 2009]

SOLUTION PROCESSED ORGANIC SEMICONDUCTOR THIN-FILM ...zc714wk4534/Stanford PhD... · semiconductor based thin-film transistors ... performance organic semiconductor materials and

  • Upload
    dodung

  • View
    228

  • Download
    0

Embed Size (px)

Citation preview

SOLUTION PROCESSED ORGANIC SEMICONDUCTOR

THIN-FILM TRANSISTORS FOR FLEXIBLE ELECTRONICS:

DEVICE PHYSICS, DEVICE MODELING, FABRICATION

TECHNOLOGY, AND INTERFACE ENGINEERING

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF ELECTRICAL ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Zihong (Bill) Liu

November 2009

[Defense Date: September 10th, 2009]

http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/zc714wk4534

© 2010 by Zihong Liu. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial 3.0 United States License.

ii

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Yoshio Nishi, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Zhenan Bao

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Krishna Saraswat

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

iv

Abstract

Organic or carbon electronics has been a fast-growing field in recent years covering a

broad range from nanoelectronic devices to macroelectronic systems. Besides the single-

graphene or single-carbon nanotube transistor toward extending the scaling limit of

traditional silicon metal-oxide-semiconductor field-effect transistor (MOSFET), organic

semiconductor based thin-film transistors have been actively investigated due to their

promise in large-area electronics fabricated on flexible substrates using low-cost

unconventional means, such as low/room-temperature printing and roll-to-roll

processing. This dissertation focuses on the study of device physics, device modeling,

fabrication technology, and interface engineering for solution-processed organic field-

effect transistors (SPOFET) for flexible electronics applications. There are primarily

four parts of contributions originated from this dissertation work.

The first part introduces the design and demonstration of high-performance, low-

voltage flexible SPOFETs fabricated on plastic substrates with a carrier mobility over

0.2 cm2/Vs, a turn-on voltage of near 0 V, and a record low subthreshold slope of ~80

mV/dec in ambient conditions. These exceptional characteristics are achieved by novel

device architecture design, 3-D statistical modeling for solution-shearing process

optimization, and phenyl-terminated self-assembled monolayer (SAM) based interface

engineering.

In the second part, SAM relevant physical effects and chemistry effects at the

organic semiconductor-dielectric interface are systematically investigated. Through

careful selection of a group of phenyl-terminated SAMs, we elucidate how the

performance and reliability of organic transistors are controlled by the critical

semiconductor-dielectric interfacial SAMs. In addition, we briefly introduce a spin-

v

coating process for depositing high-quality phenyl-terminated SAMs for organic

electronics applications.

The third part focuses on the device physics and device modeling of organic

transistors. In this dissertation work, we have proposed and developed a universal

physical model for organic transistors by incorporating both the charge injection effects

and charge transport properties, and successfully applied it to resolve many elusive

physical phenomena observed so far, such as the peculiar mobility scaling behavior with

respect to the channel length, the contact resistance effect, and the mysterious surface

potential profiles of organic transistors which have been experimentally probed yet

poorly understood. Of particular importance is that we discover an overshoot region in

the mobility scaling behavior and identified the existence of a critical channel length for

the peak field-effect mobility.

In the last part, we investigate novel contact engineering for organic transistors

toward lowering charge injection barrier and reducing the interfacial disorder width or

localization states. We have explored and demonstrated Fermi-level depinning at the

metal-organic interface for low-resistance Ohmic contacts by inserting an ultrathin

interfacial Si3N4 insulator in between. The contact behavior is successfully tuned from

rectifying to quasi-Ohmic and to tunneling by varying the Si3N4 thickness within 0-6 nm.

Detailed physical mechanisms of Fermi-level pinning/depinning responsible for the

metal-organic semiconductor contact behavior are clarified based on a proposed

lumped-dipole model.

vi

vii

To My Family Who Raised Me Up in the World of Humanity

To My Teachers Who Raised Me Up in the World of Science

viii

ix

Acknowledgments

This journey would not have been possible without the support, guidance, inspiration

and encouragement from many people to whom I am deeply and forever indebted.

First of all, I would like to extend my heartfelt gratitude to my supervisors, Professor

Yoshio Nishi in the Department of Electrical Engineering and Professor Zhenan Bao in

the Department of Chemical Engineering, who both have been rigorously advising,

strongly supporting, and generously helping me throughout my graduate studies and

research. Imagining working in both electrical engineering and chemical engineering as

handily as I do today would never come true without extraordinary guidance and

support from both of my supervisors in the two distinctive departments. Professor

Yoshio Nishi has over forty years of research and management experience in the

semiconductor industry and academia. His broad intelligence and unique insight to the

semiconductor technology have been a constant source of inspiration and

encouragement to my graduate research as well as future career. In addition, his well-

respected leadership and personality deeply influence me and set an example for us

students. Professor Zhenan Bao is a well-known expert in the field of organic

electronics and has made revolutionary contributions to the development of high

performance organic semiconductor materials and devices. Her keen sense of

experiments, grand ambitiousness of scientific research and enduring patience provided

me invaluable help to thrive in the emerging field of nano-/macro-electronics. I also

appreciate both supervisors for guiding me how to develop to a true research scientist.

I am also grateful to Professor Alberto Salleo and Professor Krishna Saraswat for

their valuable suggestions on the research and for their precious time and efforts as

members of my reading and defense committee.

x

Over the past three years of my PhD studies and research at Stanford University,

there have been constant valuable interactions and collaboration with members from

various groups focusing on semiconductor devices, organic electronics, and

nanotechnology, including Dr. Hector Becerril, Dr. Mark Roberts, Dr. Peng Wei, Dr.

Liangbing Hu, Dr. Eric Verploegen, Ajay Virkar, Dr. Toshifumi Irisawa, Dr. Joon Hak

Oh, Dr. Blanka Magyari-Kope, and Dr. Masaharu Kobayashi. I appreciate their

tremendous help and great collaboration that made this work possible. In addition, I

would like to thank Professor Yoshio Nishi’s entire group and Professor Zhenan Bao’s

entire group where numerous colleagues offered support and help along with my

research in the labs.

Special thanks go to the Toshiba Inc. who has been sponsoring this project through

Stanford CIS-FMA program. Particularly, Dr. Bipul Paul and Dr. Masaki Okajima have

been providing suggestions and discussions as the CIS-FMA project mentors, and I

appreciate their support and contributions to the success of this project.

Also, I wish to take this opportunity to extend my sincere thanks to all of my

teachers from kindergarten to graduate school who have truly raised me up in the world

of science, step by step and word by word.

Research is tough. Failures and frustrations always come along the way to the

ultimate success. I can not sustain it without mental well-being and spiritual support,

most importantly, from my family. I am indebted to my parents, my three elder sisters,

and my girlfriend for their never-fading and purely selfless love, support, encouragement

and prayers throughout my whole life. I will try my best to spend the rest of my life

making them as happy as they have made me.

xi

Contents

Abstract..................................................................................................... iv

Acknowledgments .................................................................................... ix

List of Tables........................................................................................... xv

List of Figures ....................................................................................... xvii

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

1.1 Historical Overview of Semiconductor Electronics ............................................2

1.2 Carbon Materials for Nano- and Macro-Electronics ...........................................3

1.3 Motivations and Objectives .....................................................................................5

1.4 Organization of this Dissertation............................................................................9

2 High-Performance Solution Processed Flexible Organic Transistor .. 11

2.1 Flexible SPOFET Device: Design and Fabrication .......................................... 13

2.1.1 4T-TMS Electronic Properties ............................................................................. 15

2.1.2 4T-TMS SPOFET on Silicon Substrate.............................................................. 17

2.1.3 Comparison of Different Solution Deposition Methods................................. 22

2.1.4 3-D Statistical Process Optimization................................................................... 22

2.1.5 Optimized SPOFET Performance on Silicon Substrate .................................. 33

xii

2.2 Flexible SPOFET Device: Performance and Discussion ................................. 38

2.3 Summary .................................................................................................................. 43

3 Interface Engineering by Self-Assembled Monolayer......................... 45

3.1 Background and Motivation ................................................................................. 45

3.2 Phenyl-SAM Processing and Characterization................................................... 49

3.2.1 Spin-coating Process and Mechanism ................................................................. 49

3.2.2 Spin-coated Phenyl-SAM Characterization......................................................... 52

3.3 Phenyl-SAM Effects in Organic Transistors ...................................................... 55

3.3.1 PBTTT Transistor Structure and Device Fabrication....................................... 55

3.3.2 Device Electrical Characteristics .......................................................................... 56

3.3.3 Threshold Voltage Control by SAM Dipole Moment ...................................... 61

3.4 Summary .................................................................................................................. 66

4 Device Physics and Universal Modeling of Organic Transistor ......... 67

4.1 Background and Motivation ................................................................................. 68

4.2 Universal Modeling of Organic Transistors ....................................................... 70

4.2.1 Charge Injection at the Metal-Organic Interface ............................................... 72

4.2.2 Charge Transport in the Active Layer: EPME Model ...................................... 78

4.2.3 Unified Model for Transistor Operations........................................................... 84

4.3 Simulation vs. Experimental Results.................................................................... 88

4.3.1 Transfer and Output Characteristics.................................................................... 89

4.3.2 Surface Potential Profile and Contact Resistance .............................................. 92

4.3.3 Mobility Scaling Behavior...................................................................................... 95

4.4 Summary .................................................................................................................. 99

xiii

5 Contact Engineering for Organic Electronic Device ......................... 101

5.1 Background ........................................................................................................... 101

5.2 Experimental Results on Fermi-Level Depinning........................................... 102

5.3 Theory and Proposed Model .............................................................................. 109

5.4 Summary ................................................................................................................ 114

6 Conclusions and Outlook ................................................................... 117

6.1 Conclusions ........................................................................................................... 117

6.2 Outlook.................................................................................................................. 119

Appendix A. Spin-coating Process for Phenyl-Terminated Self-

Assembled Monolayer ............................................................................ 121

Bibliography ........................................................................................... 125

xiv

xv

List of Tables

Table 2.1 Experiment design of the first 24 different processing conditions

(“BL-1”) toward systematic study and optimization of the shearing process.......24

Table 3.1 SAM molecule dipole, molecule length, and SAM density as simulated

using Gaussian package and PM3/MOPAC9 model. ..............................................64

Table 4.1 Representative parameters, corresponding physical meaning and

typical value used for the simulation based on the universal device model in this

work .................................................................................................................................88

xvi

xvii

List of Figures

Figure 1.1 The historical development of traditional silicon metal-oxide-

semiconductor field-effect transistors (MOSFET) used for VLSI circuits. Source

data: Intel and ITRS.........................................................................................................3

Figure 1.2 Schematic structure of a typical (a) bottom-gate, top-contact and

(b)bottom-gate, bottom-contact organic TFT. S: source, D: drain, G: gate. ..........5

Figure 1.3 A proposed roll-to-roll manufacturing process for solution processed

organic thin-film transistors. ..........................................................................................6

Figure 1.4 A historical overview of the organic transistor performance based on

different materials and processes. Adapted from [18]. Copyright © 2005 IEEE. .7

Figure 2.1 Illustration of the subthreshold slope (SS), as defined by

∂(logIDS)/∂VGS, which reflects how fast the transistor is switched between the on

and off state. ...................................................................................................................12

Figure 2.2 (a) Device structure and (b) photograph of our flexible 4T-TMS

SPOFETs. (c) Scanning electron microscopy (SEM) image of the bilayer PVP-

EAD dielectric that possesses apparent single-layer morphology. SEM was

performed at 3 keV in a tilt angle of 45º on an deliberately peeled-off area to

examine both the surface and the cross section. .......................................................14

Figure 2.3 (a) UV-vis absorption spectra of 4T-TMS in a solution/film phase. (b)

Cyclic voltammety (CV) diagram of 4T-TMS with a reference compound of

HOMO=-4.80 eV in 1,2-dichlorobenzene. (c) Photoelectron spectroscopy

measurement results for 4T-TMS films......................................................................16

xviii

Figure 2.4 SPOFET device structure on silicon substrate and its basic fabrication

process flow. .................................................................................................................. 18

Figure 2.5 Surface structure, water contact angle and cross-polarized microscope

image of 4T-TMS film deposited by solution shearing process on (a) PTS treated

substrate and (b) OTS treated substrate. Large crystalline domain films can be

successfully deposited on PTS substrate, while nonuniform or little film is

consistently observed on OTS substrate due to its dewettability to the 4T-TMS

organic semiconductor solutions. ............................................................................... 19

Figure 2.6 Illustration of the solution shearing process and analysis of the

shearing forces during the film growth. ..................................................................... 20

Figure 2.7 Simplified energy band diagrams without considering interface dipole

effect for the top-contact 4T-TMS SPOFET on silicon substrate along (a)

source-to-drain channel direction and (b) gate-to-film direction. The

HOMO/LUMO levels are based on the measurement as described in Section

2.1.1. ................................................................................................................................ 21

Figure 2.8 Cross-polarized optical micrographs of the 4T-TMS thin films

deposited by (a)-(b) spin-coating method, from 5 mg/mL solution in

chlorobenzene at (a) room temperature or (b) elevated temperature of ~80ºC;

(c)-(d) drop casting method (c) with or (d) without solvent annealing; (e)-(f)

solution shearing method with (e) 4T-TMS solution concentration: 6 mg/mL in

xylene, deposition temperature: 112 ºC, shearing speed: 0.10 mm/s and (f) 4T-

TMS solution concentration: 6 mg/mL in chlorobenzene, deposition

temperature: 78.6 ºC, shearing speed: 0.10 mm/s. (g) AFM tapping-mode height

image of a local film shown in (e). .............................................................................. 23

Figure 2.9 Compiled statistical analysis of 162 SPOFET samples in “BL-1”,

which shows the frequency distribution of the average mobility (μsat) with respect

to different (a) solvents; (b) solution concentrations; (c) shearing speeds; and (d)

deposition temperatures as listed in Table I. Spline interpolation algorithm is

employed to generate the plots to facilitate comparison......................................... 26

xix

Figure 2.10 Compiled statistical analysis of 162 SPOFET samples in “BL-1”,

which shows the frequency distribution of the average Ion/Ioff for different (a)

solvents; (b) solution concentrations; (c) shearing speeds; and (d) deposition

temperatures as listed in Table I. Spline interpolation algorithm is employed to

generate the plots to facilitate comparison.................................................................27

Figure 2.11 (a) The frequency distribution of μsat for the three different solvents

used in “BL-2”. Spline interpolation algorithm is employed to generate the

curves to facilitate comparison. (b)-(d) Three-dimensional surface and contour

plots of the average mobility (μsat) as a function of different shearing speeds,

solution concentrations, and deposition temperatures. Here we applied Renka-

Cline interpolation gridding algorithm and experimental data points of 172

SPOFETs in “BL-2” for the plotting. ........................................................................30

Figure 2.12 (a)-(b) Solution concentration effect, (c)-(d) shearing speed effect;

and (e)-(f) deposition temperature effect in the solution shearing process, as

revealed by the bright-field or cross-polarized microscopy images........................31

Figure 2.13 Measured film thickness on the selected wafers of “BL-1” as listed in

Table 2.1. .........................................................................................................................32

Figure 2.14 Electrical characteristics of solution-shearing processed 4T-TMS

SPOFETs with two different channel lengths (LA=35 μm and LB=16 μm), both

of which were prepared on the same wafer under the following conditions: 4T-

TMS concentration C=8 mg/mL in xylene, deposition temperature T=84 ºC,

shearing speed R=0.10 mm/s. (a) Device A shows effective mobility μsat ~0.2

cm2/Vs, Ion/Ioff ~106. (b) Device B shows μsat ~0.3 cm2/Vs, Ion/Ioff ~7×105; the

relatively poor subthreshold slope may be caused by contamination from the

short-channel shadow mask during the gold deposition..........................................34

Figure 2.15 The effective saturation mobility (μsat) and Ion/Ioff distribution of 18

measured samples fabricated on the same wafer as for devices A and B shown in

Figure 2.14. The red line is based on the Gaussian fitting model...........................35

Figure 2.16 ......................................................................................................................37

xx

Figure 2.17 Out-of-plane X-ray diffraction (XRD) spectra of solution-sheared

4T-TMS thin film. The inset is a cross-polarized optical microscopy image of the

highly crystalline film on plastic PET substrate........................................................ 39

Figure 2.18 (a) Transfer [Vds=-1.5 V] and (b) output characteristics of the

flexible 4T-TMS SPOFETs. (c) A simplified energy band diagram clarifying the

device physics responsible for the ideal turn-on voltage. (d) Device performance

with respect to different operating voltages. ............................................................. 40

Figure 2.19 A review on the representative flexible SPOFET performances

(except [54] which is based on vapor process) reported in literature in the past

decade. Our work contributes to one of the highest performances, with a

record-low subthreshold slope of ~80 mV/dec....................................................... 42

Figure 3.1 Schematic illustration of the gate voltage induced charge carrier layer,

which is close to the semiconductor-dielectric interface. ........................................ 46

Figure 3.2 Schematic illustration of the concept of dielectric surface modification

by self-assembled monolayers. .................................................................................... 47

Figure 3.3 A group of phenyl-terminated SAM molecules used for the interface

engineering study in this work. The intra alkyl chain length increases from PTS,

PETS, PBTS, to PHTS, PAPTS is comparable to PBTS except that an –NH-

group replaces a –CH2- in the intra part of PBTS. (PTS: phenyltrichlorosilane,

PETS: phenethyltrichlorosilane, PBTS: 4-phenylbutyltrichlorosilane, PHTS: 6-

phenylhexyltrichlorosilane, PAPTS: N-phenylaminopropyltricholorosilane) ...... 48

Figure 3.4 Schematic illustration of the spin-coating process for depositing

phenyl-terminated SAMs from the solution of phenyl-trichlorosilane in

anhydrous toluene. The trichlorosilane headgroup spontaneously aggregate and

align on the liquid drop surface, providing an ideal environment for the fast

reaction of the trichlorosilane with the hydroxyl group on the substrate surface

[70-71]. The anhydrous toluene solvent being selected here is critical to the

xxi

process. The conjugated aromatic group of toluene is supposed to interact with

the functional endgroup of the phenyl-silane and facilitate the alignment of the

phenyl-silane on the liquid drop surface.....................................................................51

Figure 3.5 A preliminary screening of different film morphology as deposited by

different spin-coating processes for PBTS. Images were taken under bright-field

microscope. The solvent, concentration and spin rate/time are all critical to high

quality phenyl-SAM. ......................................................................................................51

Figure 3.6 Measured water contact angle on phenyl-SAM modified SiO2

surfaces. The molecular length of each SAM molecule is computed by

PM3/MOPAC9 model and shown for comparison purpose here.........................53

Figure 3.7 Representative phenyl-SAM film morphology for (a)

PTS/PETS/PBTS/PHTS and (b) PAPTS, as deposited by the optimized spin-

coating process. For PTS-series SAM, the typical roughness is 0.1-0.2 nm (rms);

for PAPTS SAM, the roughness is 0.2-0.3 nm (rms). This indicates the spin

coating process yields excellent phenyl-SAM quality................................................53

Figure 3.8 GIXD image of the PHTS SAM on native silicon oxide surface as

deposited by the spin coating process. By courtesy: Dr. Eric Verploegen............54

Figure 3.9 Measured phenyl-SAM thickness (d) using ellipsometry and the

simulated molecule length (L) based on PM3/MOPAC9 model or Gaussian

Package 03’, B3LYP/6-31G(d) model. The average title angle of the SAM

molecule with respect to surface normal direction, θ, is estimated by cos

(θ)=d/L, where L is based on the Gaussian simulation here. .................................55

Figure 3.10 Device structure of PBTTT transistor incorporating different

phenyl-SAMs at the dielectric-semiconductor interface. Courtesy: PBTTT was

provided by Dr. Iain McCulloch and Dr. Martin Heeney (formerly in Merck)....56

Figure 3.11 Representative (a) transfer and (b) output curves for the PBTTT

transistors based on the spin-coated phenyl-SAMs. The results shown here are

specifically based on PBTS...........................................................................................57

xxii

Figure 3.12 (a) Mobility and (b) threshold voltage of PBTTT transistors

incorporated with different phenyl-terminated SAMs at the dielectric-

semiconductor interface. .............................................................................................. 58

Figure 3.13 General mechanisms responsible for the hysteresis effect in the

organic transistor. Only (1) the charge injection from semiconductor to the

interface induces counterclockwise hysteresis in the transfer curve. The other

mechanisms including (2) charge injection from the gate electrode, (3) residual

dipole caused by slow polarization in the bulk dielectric, and (4) mobile ions in

the dielectric contribute to the counterclockwise hysteresis in the transfer curve.59

Figure 3.14 General hysteresis effect for PBTTT transistors modified with

different phenyl-SAMs. PTS/PETS/PBTS/PHTS show no appreciable

difference in the hysteresis, therefore only a representative I-V hysteresis curve

is given here as adopted from PBTS based transistors............................................ 60

Figure 3.15 Bias stress effect on the PBTTT transistors with different phenyl-

SAMs. Considerably larger bias stress effect is observed for the PAPTS modified

interface. The applied gate bias for the bias stress measurement is VGS=-60V.

For each I-V measurement during the bias stress test, a recovery time of 200

seconds is given to allow the fast reversible traps recovered.................................. 61

Figure 3.16 Simplified energy band diagram of the PBTTT transistor

incorporating the SAM modification layer at the dielectric-semiconductor

interface. The SAM dipole creates an additional electric field normal to the

interface, bending the energy alignment and thus tuning the threshold voltage. 62

Figure 3.17 A simple model to describe the dipole moment effect in tuning the

surface charge density and transistor’s threshold voltage. The SAM dipole (p) on

the dielectric surface induces additional surface charge (QS). Θ is the tilt angle of

the dipole moment. ....................................................................................................... 62

Figure 3.18 Measurement results vs modeling results for the threshold voltage

shift with respect to the surface SAM dipole and SAM induced surface charge

xxiii

density. Both p-type PBTTT transistors and n-type PCBM transistors are

measured in this work. ..................................................................................................65

Figure 3.19 Schematic illustration of the proposed electric shielding tensors

effect of the SAM at the dielectric-semiconductor interface: strong external

interfacial electric field as induced by the gate voltage would distort the inherent

dipole moment of the SAM molecule, thus influencing the performance and

reliability of the organic transistors. ............................................................................66

Figure 4.1 A generalized 1-D representation of the physical process dominating

the device operation for organic transistors...............................................................70

Figure 4.2 The energy level diagram of the metal-organic semiconductor

interface. ..........................................................................................................................73

Figure 4.3 The calculated boundary condition for diffusion-limited charge

injection at metal-organic semiconductor interface. .................................................75

Figure 4.4 (a) Schematic illustration of the transition region close to the contact

and the channel region in the organic semiconductor film; (b) The DOS profiles

of the transition region and the major channel based on the mobility edge

model; (c) An equivalent representation of (b) with the critical energy for both

regions set to be identical to simplify the mathematical derivation........................77

Figure 4.5 The relationship between the tail width of localized states (∆Etail), the

relative Fermi energy level (Ef-E0), and the applied gate voltage (VGS). The

following conditions are employed for the calculation here: 300 nm SiO2 gate

dielectric, Von=0 V, h=1 nm, Ntail=1021/cm3, Ec=40 meV, T=300 K. ...................81

Figure 4.6 Simulation results of the (a) output and (b) transfer characteristics,

based on the universal device model introduced in Section 4.2. In addition to

parameters given in Table 4.1, the followings are included: L=10 µm, γ(tr)(1/Tn)=

0.001 cm1/2V-1/2, γ(ch)(1/Tn)=0, (a) φb=0.7 or 0.2 eV, σ=0.1 eV, and (b) φb=0.7 eV,

xxiv

σ=0.1 eV. Note: for the transfer curve simulation, an Ioff (IDS@VGS=0)=1 pA is

assumed per practical situations. ................................................................................. 91

Figure 4.7 (a) Our simulated surface potential profile of a representative organic

transistor operating in different conditions, either in linear regime or saturation

regime, upon the applied drain voltage, VDS, for a fixed gate voltage of VGS=-

100 V. Here T=300 K. (b) A representative surface potential profile of the

P3HT polymer transistor, as recorded in experiments by noncontact

potentiometry using scanning Kelvin probe microscope by Bürgi et al. [112].,

copyright © American Institute of Physics (AIP) 2003........................................... 93

Figure 4.8 Simulated surface potential profiles along the channel direction for

the same transistor only with different injection barrier height (φb=0.7 eV or 0.3

eV) and localized trap states tail width (σ=100 meV or 50 meV). The inset is a

magnified view in the 200 nm region close to the source contact. ........................ 94

Figure 4.9 Mobility scaling behavior with respect to different physical origins

including carrier injection barrier height (φb), localized trap states energy

distribution (σ), Poole-Frenkel like field dependence of carrier mobility (γ(1/Tn)),

and channel length dependent film crystallinity (γc). For simplicity of the

simulation here, the film crystallinity effect is lumped into the mobility’s electric

field effect. ...................................................................................................................... 98

Figure 4.10 Measurement results of the extrinsic field-effect mobility with

respect to different channels for the solution-shearing processed polycrystalline

4T-TMS SPOFETs. The device fabrication and measurement are detailed in

Chapter 2. All of these devices are fabricated on the same wafer with the same

process. The dashed line is for visual guidance and refers to the modeling and

simulation results shown in Figure 4.9. It is found that the finding based on our

model is in excellent agreement with the experimental results............................... 99

xxv

Figure 5.1 Device structure and process flow of the Au/Si3N4/pentacene diodes.

Pentacene was chosen for this study since the pinning factor (S=dφb/dψmetal)

S~0.4 is relatively small for the metal/pentacene interface...................................103

Figure 5.2 AFM images of 200 nm pentacene films deposited on Au at room

temperature (a) before and (b) after Si3N4 sputtering deposition. Height scale:

300 nm...........................................................................................................................104

Figure 5.3 I-V measurement results showing the diodes with/without Si3N4

based on evaporated Au pads are regularly shorted due to metal atoms

penetration along the grain boundaries or the roughness-induced vacancies.....104

Figure 5.4 I-V characteristics of the Au/Si3N4/pentacene diodes with different

Si3N4 thicknesses (with Au wire as the top cathode electrode), providing direct

evidence that the Au/pentacene diode has been successfully tuned to rectifying,

quasi-Ohmic, and symmetric tunneling behavior by modulating the Si3N4

thickness. (The effective contact area between Au wire and Si3N4/pentacene is

only a fraction of the wire cross section due to the surface roughness of

Si3N4/pentacene layer, and may vary from device to device. One should note

that the I-V shape giving the diode property is the essence here.).......................106

Figure 5.5 Normalized dynamic resistance (RAC=∂V/∂I) and static resistance

(RDC=V/I) of the Au/Si3N4/pentacene diodes with respect to the Si3N4

thickness, as calculated from their respective I-V measurement curves.

Normalization is based on the fact that the maximum forward-biased diode

current (here at V=+1 V) is less affected by the Si3N4 thickness. Note: The

device with a thick Si3N4 layer (t=6.1 nm) was not normalized here as tunneling

dominates therein.........................................................................................................107

Figure 5.6 Device structure, microscopy image, and process flow of the

Ag/Si3N4/PTCDA diodes. PTCDA was chosen here since the pinning factor

(S=dφb/dψmetal) is S~0 for metal/PTCDA interface, indicating very strong Fermi-

level pinning effect at the metal/PTCDA interface................................................108

xxvi

Figure 5.7 (a) A quick and straightforward extraction method for the relative

contact resistance of the diodes with different Si3N4 thicknesses, through

measurement of the I-V characteristics on two adjacent electrodes; (b) A

minimum contact resistance is observed at the Ag/PTCDA interface with a

sandwiched Si3N4 thickness of ~2.5 nm, reaffirming the Fermi-level depinning at

M/O interfaces by inserting the ultrathin interfacial Si3N4 layer.......................... 109

Figure 5.8 Different mechanisms of Fermi-level pinning at the metal/organic

semiconductor interface, including (a) MIGS and intrinsic surface states, (b)

charge transfer, (c) covalent bonding, (d) permanent molecular dipole, (e) image

force, and (f) exchange/Pauli repulsion. An interface dipole is always created

upon the M/O junction formation........................................................................... 110

Figure 5.9 M/O interface energy band diagrams based on the proposed lumped

interface dipole model: (a) before M/O interface formation; (b) upon M/O

interface formation, Fermi-level pinning arises from various interface dipole

elements; (c) after inserting an ultrathin Si3N4 insulator, the Fermi-level

depinning takes effect by blocking the physisorption/chemisorption................ 112

Figure 5.10 Simulation results based on the proposed simple depinning model

showing the dominant mechanisms responsible for the contact resistance

behavior after insertion of the ultrathin insulating layer at the M/O interface. 114

1

Chapter 1

Introduction

We choose to go to the moon. We choose to go to the moon in this decade and do the other

things, not because they are easy, but because they are hard, because that goal will serve to

organize and measure the best of our energies and skills, because that challenge is one that we

are willing to accept, one we are unwilling to postpone, and one which we intend to win.

- John F. Kennedy, 1962

Imagine a world filled with flexible personal digital assistants (PDA) or cellphones,

large-area rollable displays, bendable scanners, wearable electronic clothes, as well as

many other electronic devices with versatile functions, fantastic appearance and

mechanical flexibility. An intuitive question would be if there is any possibility to

integrate the well-established semiconductor device technology traditionally based on

silicon hard materials into these novel applications. The prospect seems optimistic, yet,

significant challenges remain both in scientific understanding of the newly emerging

phenomena and in engineering of the reliable devices.

2 CHAPTER 1. INTRODUCTION

1.1 Historical Overview of Semiconductor Electronics

It is natural to learn from history since history always guides future. Along with the

development of semiconductor electronics over the past few decades, people have been

broadly exploring two intriguing fields: How to make the devices extremely small?

And, how to make the systems extremely large?

Specifically, the first one is toward the limit of scaling down each component,

namely for “nanoelectronics” nowadays, and the other one is toward the limit of scaling

up the system, namely for “macroelectronics”. Identifying their specific applications and

understanding their unique characteristics are required for the success in engineering

devices and systems toward both extremes. To be accurate while without loss of

generality, we’ll focus on discussing a most representative and impotant applications for

each area, i.e. very large scale integration (VLSI) circuit chips and large area displays,

respectively. Figure 1.1 shows the historical development of traditional silicon metal-

oxide-semiconductor field-effect transistors (MOSFET) which is the central component

of VLSI chips. The primary focuses in such nanoelectronics area include reducing the

cost per function or cost per transistor, scaling down of the device to smaller feature

size, packing higher integration density, increasing circuit operation speed, and

minimizing chip power dissipation. The future, however, is challenged by the scaling

bottleneck beyond current 15/22 nm feature size due to fundamental physical effects

arising from materials, processes, and the device architectures [1].

In contrast, for macroelectronic applications as represented by displays, efforts are

centered on reducing the cost per area in addition to expanding the system to larger size

with increased resolution, higher mechanical flexibility, enhanced robustness, and lighter

weight. The requirements of the dimensional scale and operation frequency on its core

component for the backplane, typically hydrogenated amorphous silicon (α-H:Si) or

polycrystalline silicon thin-film transistors (TFT), are less stringent than the MOSFETs

used for high-performance VLSI circuits. The future of macroelectronic applications

relies more heavily on the cost per area and ubiquity limit opposed by traditional

materials, processes, and devices employed in the large-area electronics.

CHAPTER 1. INTRODUCTION 3

1970 1980 1990 2000 2010 202010-9

10-7

10-5

10-3

10-1

101

Year

Cos

t per

Tra

nsis

tor (

$)

100

101

102

103

104 Minim

um Feature S

ize (nm)

Figure 1.1 The historical development of traditional silicon metal-oxide-semiconductor field-effect transistors (MOSFET) used for VLSI circuits. Source data: Intel and ITRS.1

1.2 Carbon Materials for Nano- and Macro-Electronics

In order to circumvent the aforementioned limits, there has been extensive research

work in recent years on exploring novel semiconductor materials and their relevant

processes and devices. Among many others, carbon (organic) electronic materials are

found very promising in both nanoelectronics and macroelectronics areas. Carbon

electronics has emerged as a fast-growing field covering a broad range from

nanoelectronic devices to macroelectronic systems.

On one hand, individual carbon nanotube, graphene, organic molecule, or carbon

nanoribbon was reported having superior electrical properties with extraordinary

intrinsic carrier mobility over 100,000 cm2/Vs, up to 1,000 times higher than in bulk

silicon [2-3]. Consequently, single-graphene [4] or single-carbon nanotube transistor [5]

provides a possible route toward extending the scaling limit of traditional silicon

1 Intel and Dataquest reports (December 2002), see Gordon E. Moore, “Our Revolution”, http://www.sia-online.org/galleries/default-file/Moore.pdf; International Technology Roadmap for Semiconductors (ITRS), http://www.itrs.net/

4 CHAPTER 1. INTRODUCTION

MOSFET. The main hurdle at present is the poor reproducibility of identical devices

required for large scale integrated circuits. How to precisely control the device

placement and integration at will is a major challenge on pushing forward the carbon

nanomaterials into high performance nanoelectronic applications.

On the other hand, carbon nanotube, small-molecule and polymer organic

semiconductor thin-film1 transistors can be readily fabricated with large-scale control of

positioning and integration, though the electrical performance and dimensional scaling

capability is more confined. Typically, the carrier mobility in these thin films is on the

order of 0.1-10 cm2/Vs [6], two to three orders lower than that in single crystal silicon,

thus being inappropriate for high speed computing where traditional single crystal

silicon still dominates. Fortunately, such performance still approaches, if not surpasses,

that of hydrogenated amorphous silicon (α-Si:H) TFTs which are widely used in display

backplanes [7]. Moreover, their unique possibility of low-temperature deposition and

excellent mechanical flexibility enable us to fabricate large-area electronics on flexible,

light-weight substrates using low-cost unconventional means, such as room-temperature

printing and roll-to-roll processing [8]. Therefore, carbon semiconductor based thin film

devices are highly promising for novel and practical macroelectronic applications where

relatively low mobility 2 and medium device feature size satisfy the requirements.

Foreseeable examples include electronic paper, flexible displays, ubiquitous electronic

walls, organic radio-frequency identification (RFID) tags, organic solar cell and light-

emitting diode arrays, chemical sensors, and e-clothes.

For this dissertation work, we focus on studying the organic semiconductor based

thin-film transistors for the aforementioned macroelectronics and flexible electronics

applications.

1 Thin film typically consists of arrays or a number of individual molecules or nanotubes that form a thin layer of film. 2 In comparison to single crystal silicon or individual carbon nanotube, graphene, and nanoribbon for high performance nanoelectronic devices.

CHAPTER 1. INTRODUCTION 5

1.3 Motivations and Objectives

As discussed in Section 1.2, organic semiconductor based thin-film transistors have

attracted wide attention in recent years due to their promise in low-cost, large-area,

flexible electronics and the compatibility with scalable fabrication process. Figure 1.2

illustrates the typical structure of a bottom gate, top contact and bottom gate, bottom

contact organic TFT. Besides conducting materials for the source (S), drain (D), and

gate (G), dielectrics for isolating the channel from the gate and organic semiconductor

thin films acting as the active channel are also crucial to the organic TFT operation.

GDielectric

Organic Semiconductor Thin FilmS D

GDielectric

S Organic Semiconductor D

Figure 1.2 Schematic structure of a typical (a) bottom-gate, top-contact and (b)bottom-gate, bottom-contact organic TFT. S: source, D: drain, G: gate.

Of particular interest in the device fabrication is how to deposit the active channel

materials. In general, organic semiconductor thin films can be either vapor processed or

solution processed, depending on their physical and chemical properties. Vacuum

evaporation as a mature fabrication technology is widely used for most organic small-

molecule semiconductors to grow highly uniform films on the substrate. This is

particularly important for large area circuits and systems. However, vapor process

suffers from relatively high temperature in evaporation, high fabrication cost from

expensive setup and material waste, and low throughput as confined by the equipment

space.

An alternative to the vacuum evaporation deposition method is solution processing.

The concept is simple: when the materials are soluble to some specific solvents, one can

6 CHAPTER 1. INTRODUCTION

deposit the film by drop casting, spin-coating, or other ways from the material’s solution

phase [9]. Advantages of solution processing of organic semiconductors include:

(1) It is compatible with printing techniques toward printed electronics1. Since

the devices can be fabricated with screen printing [10], inkjet printing [11-13],

or microcontact printing [14], they possess benefits from all these well-

developed existing processes.

(2) It is possible to manufacture the devices on flexible substrates using a fast,

roll-to-roll process. Figure 1.3 shows a conceptual roll-to-roll manufacturing

process for solution processed organic transistors. The flexible substrate

moves between two end rollers; modules (1)-(n) serve different layer

depositions. Configurations for each module include three substrate-

supporting transfer rollers, a shearing roller clamping the substrate with

designed pressure, and an organic solution tank close to the shearing roller.

Figure 1.3 A proposed roll-to-roll manufacturing process for solution processed organic thin-film transistors.

1 See http://flextech.org/.

CHAPTER 1. INTRODUCTION 7

(3) Low temperature processing becomes available. In solution phase deposition,

the temperature is typically below 100ºC or even down to room temperature,

being significantly lower than that required in vapor process [15].

(4) There is no throughput limit against evaporation setup space.

(5) Self-assembly in solution phase provides a novel method for versatile

patterning of the devices [16-17].

(6) The fabrication cost is considerably reduced based on the above superiority.

Despite the great potential of solution processed organic thin-film field-effect

transistors (SPOFET), their electrical performance as show in Figure 1.4 is still lagging

behind the counterparts based on vapor process due to unfavorable film microstructures

and poor uniformity. A number of issues remain for SPOFET although the

performance has been improved dramatically in the past decade.

Figure 1.4 A historical overview of the organic transistor performance based on different materials and processes. Adapted from [18]. Copyright © 2005 IEEE.

8 CHAPTER 1. INTRODUCTION

First, it is challenging to fabricate uniform and highly crystalline films via solution

deposition of small-molecule organic semiconductors, albeit many novel techniques

have been proposed [9]. Polymer semiconductors can be readily spin coated from

solution phase to yield uniform films, but the mobility in general is inferior to that of

small molecules as shown in Figure 1.4. Finding an efficient and optimized solution

process is thus important, especially for small-molecule organic semiconductor based

transistors.

Secondly, SPOFETs fabricated on flexible substrates normally show lower

performance as compared to their counterparts on rigid substrates such as on silicon or

glass. It is essential to achieve high performance SPOFETs on flexible substrates,

including both high mobility and steep subthreshold slope, toward practical flexible

electronics applications.

Additionally, the device physics of organic thin-film field-effect transistors,

especially SPOFETs, remains elusive in many aspects [19]. For instance, further

fundamental understanding and studies are in demand for charge transport properties in

the organic semiconductor films, change injection or contact properties at the metal-

organic semiconductor interface, and interface properties between the organic

semiconductor and the dielectric, all of which are critical in determining the device

performance and device engineering protocols.

Lastly, unlike that for traditional silicon MOSFET where industrial standard device

modeling and simulation programs have been well established1, modeling and simulation

for organic transistors remain as a challenging task. Indeed, a device model

incorporating the fundamental physical effects must also help resolve many elusive

phenomena observed from the electrical characterization.

In this dissertation, we address all the above issues and focus on the study of device

physics, device modeling, fabrication technology, and interface engineering for solution-

processed organic field-effect transistors for flexible electronics applications.

1 For instance, TCAD Sentaurus from Synopsys Inc., http://www.synopsys.com/tools/tcad

CHAPTER 1. INTRODUCTION 9

1.4 Organization of this Dissertation

The remainder of this dissertation is organized as follows.

Chapter 2 presents the design and demonstration of high-performance, low-voltage

flexible SPOFETs fabricated on plastic substrates with a carrier mobility over 0.2

cm2/Vs, a turn-on voltage of near 0 V, and a record low subthreshold slope of ~80

mV/dec in ambient conditions. These exceptional characteristics are achieved by novel

device architecture design, 3-D statistical modeling for solution-shearing process

optimization, and phenyl-terminated self-assembled monolayer (SAM) based interface

engineering.

In Chapter 3, we systematically investigate dipole moment related physical effects

and chemistry effects of the SAM at the organic semiconductor-dielectric interface.

Through careful selection of a group of phenyl-terminated SAMs, we elucidate how the

performance and reliability of organic transistors are controlled by their interface

conditions. In addition, we briefly introduce a simple spin-coating process for depositing

high-quality phenyl-terminated SAMs for organic electronics applications.

Chapter 4 focuses on the device physics and device modeling of organic transistors.

We have developed a universal physical model for organic transistors by incorporating

both charge injection effects at the metal-organic interface and charge transport

properties in the organic semiconductor film, and successfully applied the device model

to resolve many elusive physical phenomena observed so far, such as the peculiar

mobility scaling behavior, the contact resistance effect, and the mysterious surface

potential profiles along the channel which have been experimentally probed yet poorly

understood. The complete model as derived in an analytical manner and the comparison

between simulation results and experimental results are given in detail. Of particular

importance is that we discover an overshoot region in the mobility scaling behavior and

identified the existence of a critical channel length for the peak field-effect mobility.

To the end of lowering charge injection barrier and reducing the interfacial disorder

width or localization states, in Chapter 5, we explore and demonstrate Fermi-level

10 CHAPTER 1. INTRODUCTION

depinning at the metal-organic interface for low-resistance Ohmic contacts by inserting

an ultrathin interfacial Si3N4 insulator in between. The contact behavior is successfully

tuned from rectifying to quasi-Ohmic and to tunneling by varying the Si3N4 thickness

within 0-6 nm. Detailed physical mechanisms of Fermi-level pinning/depinning

responsible for the metal-organic semiconductor contact behavior are clarified based on

a proposed lumped-dipole model.

We discuss the future work and conclude this dissertation in Chapter 6.

11

Chapter 2

High-Performance Solution

Processed Flexible Organic

Transistor

Logic will get you from A to B. Imagination will take you everywhere.

- Albert Einstein

As outlined in the previous chapter, there are increasing research efforts focusing on

solution-processed organic field-effect transistors (SPOFET) due to their promise in

large-area electronics fabricated on flexible substrates using low-cost unconventional

means, such as low/room-temperature printing and roll-to-roll processing [6, 9].

Pioneering works on flexible, solution-processed organic transistors in the past decade have

demonstrated many intriguing applications on plastic substrates, including all-polymer

integrated circuits [12, 20], flexible smart pixels [12, 21], plastic sensors [22], rollable

1 Note: Portions of this chapter are reproduced, with permission, from [15] © 2009 IEEE, [36] © 2009 American Institute of Physics and [37] © 2008 IEEE.

12 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

displays [14, 23-24], and RFID tags [13, 25-26]. Compared to their counterparts on rigid

substrates like on glass or silicon wafer, flexible SPOFETs normally exhibit lower

performance. Typically, the charge carrier mobility is on the order of 10-2-10-1 cm2/Vs

and the subthreshold slope (SS), as illustrated in Figure 2.1, is either over 1 V/dec or

being overlooked although it is a critical parameter in determining the device switching

speed.

1log( ) ln10DS

GS

I kTSSV e

−∂= > ⋅

Figure 2.1 Illustration of the subthreshold slope (SS), as defined by ∂(logIDS)/∂VGS, which reflects how fast the transistor is switched between the on and off state.

More recently, flexible SPOFETs with a mobility over 0.1 cm2/Vs [13, 24-30] or a

rarely low subthreshold slope down to 100 mV/dec [29] have been demonstrated by

virtue of improved processing methods, a series of newly discovered high-mobility and

soluble organic semiconductors, and novel dielectric materials [29]. Nevertheless,

obstacles in achieving both specifications simultaneously still appear [31], especially for

flexible SPOFETs based on common polymeric dielectrics such as photoresist [20],

polyimide [21, 32], poly(vinyl alcohol) [33], poly(methylmethacrylate) [30], spin-on

silsesquioxane glasses [14, 34], and poly(4-vinylphenol) (PVP) [12, 23, 25-26, 28]. In fact,

these polymeric dielectrics are highly favorable in flexible electronics due to their

simplicity in synthesis, low cost, solution processability and their match of thermal

expansion coefficient with plastic substrates. It is therefore crucial to improve both the

mobility and subthreshold slope in parallel for flexible SPOFETs with polymeric

dielectrics.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 13

In this chapter, we address the above issue through a combination of the device

architecture design, solution process optimization, organic semiconductor/polymeric

dielectric development [35] and their interface engineering. Flexible SPOFETs

fabricated and tested in ambient conditions on rough plastic with a high mobility over

0.2 cm2/Vs, a turn-on voltage of near 0 V, and a subthreshold slope of ~80 mV/dec

have been successfully demonstrated [36].

2.1 Flexible SPOFET Device: Design and Fabrication

We emphasize that both the mobility and the subthreshold slope of organic transistors

rely on the semiconductor film crystallinity and the interface conditions. Therefore,

priorities for the design and fabrication of our flexible SPOFET devices are minimizing

interface and bulk trap states and enhancing the organic film crystallinity.

Figure 2.2 (a) and (b) show the device structure and a photograph of our flexible

SPOFETs, respectively. For this work, a transparent indium-tin-oxide (ITO) coated

polyethylene terephthalate (PET, 175 μm) was selected as the device substrate from its

prevalent use in organic light-emitting diodes (OLED), thus providing the possibility of

integration of OFETs and OLEDs together for all-organic flexible displays. After brief

treatment in an oxygen plasma, the ITO/PET substrate was coated with a conducting

polymer, poly(3,4-ethylenedioxythiophene) poly(styrenesulfonate) (PEDOT:HAPSS)

[Agfa-Gevaert], as the gate electrode via spin coating and cured at 100 ºC for 1 hr,

yielding a 30-50 nm film with surface roughness (rms) reduced from 3.5-9.8 nm to ~2-3

nm. The subsequent dielectric is a recently developed low-temperature cross-linkable

PVP matrix with high stability in ambient air [see Figure 2.2 (a)] [35].

Ethylenediaminetetraacetic dianhydride (EAD) [Sigma-Aldrich] is used as the cross-

linker here. Due to the original rough substrate surface and sporadic spikes, we used a

bilayer of PVP-EAD to reduce gate leakage induced by through pinholes. The first layer

of insulating film was deposited by spin coating from 105 mg/mL solution of

PVP:EAD (20:1) in propylene glycol monomethyl ether acetate (PGMEA):N,N-

dimethylformide (DMF) (3:1) at 7 krpm and cured in a vaccum oven at 90-100 ºC for 2

14 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

hr to promote the cross-linking reaction. The second layer was spin-coated from a lower

concentration of PVP:EAD solution of 63 mg/mL at 5 krpm to better tune the surface

morphology and eliminate pinholes in the dielectric, and again followed by annealing.

Bilayer PVP-EADPEDOT:PSSITO

500 nm

ITO

PET(Polyethylene-Terephthalate)

Bilayer Cross-linked PVP (PVP-EAD)

Au

PEDOT:PSS

Au

SS S

S Si

175 μm

4T-TMSPBTS-SAM

N

O

O

n

OHO

N

O

O

n

OOH

SiO O

O

SiO

O

SiO

O

SiO

O

(a)

(b) (c)

Figure 2.2 (a) Device structure and (b) photograph of our flexible 4T-TMS SPOFETs. (c) Scanning electron microscopy (SEM) image of the bilayer PVP-EAD dielectric that possesses apparent single-layer morphology. SEM was performed at 3 keV in a tilt angle of 45º on an deliberately peeled-off area to examine both the surface and the cross section.

As shown in Figure 2.2 (c), our bilayer PVP-EAD dielectric of ~180 nm possesses

apparent single-layer morphology without distinguishable traces in between or on the

top surface, as revealed by scanning electron microscopy (SEM) [FEI XL30 Sirion]. The

device yield based on this bilayer PVP-EAD was found 50-60% higher than that of

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 15

single layer dielectric. Typical surface roughness value of the bilayer PVP-EAD dielectric

is ~0.8-1.3 nm (rms).

Since phenyl-terminated self-assembled monolayers (SAM) have been found

advantageous in solution processing of organic transistors due to their decent tradeoff

between wettability and surface energy [15] as detailed in Section 2.1.2 and Chapter 3,

here, we applied 4-phenylbutyltrichlorosilane (PBTS) to modify the PVP-EAD dielectric

surface via spin coating, resulting in a high water contact angle of ~86ºwhile still

maintaining excellent wettability for organic semiconductor solutions. Systematic study

of the phenyl-terminated SAM effects will be presented in Chapter 3.

A subsequent critical step in fabricating these flexible SPOFETs is the deposition of

the organic semiconductor thin films. Here we used a recently synthesized p-type small-

molecule organic semiconductor, trimethyl-[2,2';5',2'';5'',2'''] quarter-thiophen-5-yl-silane

(4T-TMS) [the synthesis details reported by Dr. Mark Roberts in [35]; chemical structure

shown in Figure 2.2 (a)]. 4T-TMS is an air-stable and solution-processable organic

material which shows relatively high carrier mobility. To deposit 4T-TMS thin films as

the transistor’s active layer, we introduce a solution-shearing process that has been

demonstrated with capability of fast screening small-molecule semiconductor

performance [37-39]. The details of the material characterization and the 3-D statistical

process optimization will be given in the following section.

To complete the top-contact flexible 4T-TMS SPOFETs shown in Figure 2.2, 40

nm gold source and drain electrodes were thermally evaporated onto the 4T-TMS films

through a shadow mask at a rate of ~0.5 Å/s. Material Characterization and Process

Optimization

2.1.1 4T-TMS Electronic Properties

Rational design of an electronic device requires understanding of the physical and

electronic properties of its materials as well as the device energy band diagrams. Figure

2.3 (a) shows the UV-vis absorption spectra of the 4T-TMS solution in p-xylene and its

film on glass, indicating a bandgap (Eg) of 2.69 eV for the 4T-TMS molecule in the

16 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

300 400 500 600 700 800

0.00

0.25

0.50

0.75

1.00

Abs

orpt

ion

(a.u

.)

Wavelength (nm)

4T-TMS in solution

4T-TMS in film

-2 -1 0 1

-3

-2

-1

0

1

2

3

Oxidation/Reductionpeak of 4T-TMSmolecule

Cur

rent

(μA

)

Potential (V)

Oxidation/Reduction peak of reference molecule with HOMO=-4.80eV

(a)

(b)

(c)

4.6 4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2

0

1000

2000

3000

4000

5000

Yie

ld1/

2 (cps

1/2 )

Yield (cps) Yield1/2 (cps1/2)

Energy (eV)

Yie

ld (c

ps)

0

10

20

30

40

50

60

70

Figure 2.3 (a) UV-vis absorption spectra of 4T-TMS in a solution/film phase. (b) Cyclic voltammety (CV) diagram of 4T-TMS with a reference compound of HOMO=-4.80 eV in 1,2-dichlorobenzene. (c) Photoelectron spectroscopy measurement results for 4T-TMS films.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 17

solution as estimated from the absorption onset wavelength (λonset=462 nm), and a close

Eg of 2.54 eV for the 4T-TMS film on glass. The highest occupied molecular orbital

(HOMO) energy level of the 4T-TMS molecule was measured using cyclic voltammetry

(CV) [CH Instruments, Inc.] as shown in Figure 2.3 (b) and was estimated from the

onset potential [40] to be around -5.27 eV. The lowest unoccupied molecular orbital

(LUMO) energy level thus can be calculated by ELUMO=EHOMO+Eg=-2.58/-2.73 eV.

Photoelectron spectroscopy (PES) spectra [Model AC-2, Riken Keiki Co.] in Figure 2.3

(c) show the ionization potential of the 4T-TMS thin film is approximately 5.20 eV, in

good agreement with the above electrochemical measurement results.

2.1.2 4T-TMS SPOFET on Silicon Substrate

For the purpose of material characterization and solution process development/

optimization, we first fabricated top-contact bottom-gated SPOFETs on silicon

substrate instead of on flexible plastic due to their ease of processing and the operation

reliability. The device structure and its basic process flow are shown in Figure 2.4. All

fabrication and characterization of these SPOFETs were carried out under ambient

conditions with exposure to air and room light except for the Au electrode deposition.

Detailed fabrication process is as follows.

Highly doped n-type (100) silicon wafers (R≈0.0015-0.007 Ωcm) with a 310 nm

thermally grown dry oxide gate dielectric film were used as the device substrates. After

cleaning in a piranha solution (highly corrosive and oxidizing 7:3 mixture of H2SO4 and

H2O2), the wafers were further treated in a UV ozone cleaner for 10-15 minutes to

ensure removal of any organic contamination and to facilitate the subsequent surface

treatment. The silanol groups on the oxide surface were passivated with

phenyltriethoxysilane (PTS) [Aldrich Chem. Co., see Fig. 1] by immersing the wafers in

an HCl-acidified PTS solution in toluene (100 mL toluene, 3 mL PTS and 2 drops of

HCl) for over 12 hrs, then sonicated and rinsed with toluene, acetone and isopropanol,

and finally dried with nitrogen stream.

18 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

Phenyl-Triethoxy-Silane (PTS)

SiO

O O

SS S

S Si

4T-TMS

Thermally grown SiO2

N++ Si (Gate)

4T-TMS Active LayerD

Phenyl-SAM

S4T-TMS Active Layer

D S

S

S

S

SSi

S

S

S

SSiAu Au

1 Piranha cleaning for the silicon oxide wafer

2 Surface pre-treatment in UV ozone cleaner

3 PTS SAM deposition on the SiO2 surface

5 Solution deposition of 4T-TMS film

6 Annealing in vaccum oven at 80ºC for overnight

7 Thermal evaporation of the Au electrode for S/D

4 Sonication cleaning for the surface, soaked in toluene

Figure 2.4 SPOFET device structure on silicon substrate and its basic fabrication process flow.

As shown in Figure 2.5, the water contact angle for the PTS-modified wafer is ~70º.

4T-TMS film is readily deposited on the PTS-modified wafer surface by the solution

shearing method described below, while non-uniform or little film is consistently found

with our deposition on the wafers modified by n-octadecyltrichlorosilane (OTS)

[Aldrich Chem. Co.] SAM which is widely used for vapor-processed organic transistors

[41-42], primarily due to the high hydrophobicity of OTS with water contact angle of

~107º and its dewettability to the organic semiconductor solutions. This observation

indicates the importance of the aromatic group in PTS in facilitating wettability and the

film deposition for SPOFETs, and is in obvious contrast to traditional argument that

lower surface energy for surface with larger water contact angle (CA) usually leads to

higher performance of organic transistors.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 19

Si

O

O

O

Si

O

O

Si

O

O

Si

O

O

OTS

CA~107O

(a)

(b)

shea

ring

100 µm

Si

O

O

O

Si

O

O

Si

O

O

Si

O

O

PTS

CA~70O

Figure 2.5 Surface structure, water contact angle and cross-polarized microscope image of 4T-TMS film deposited by solution shearing process on (a) PTS treated substrate and (b) OTS treated substrate. Large crystalline domain films can be successfully deposited on PTS substrate, while nonuniform or little film is consistently observed on OTS substrate due to its dewettability to the 4T-TMS organic semiconductor solutions.

4T-TMS solutions are prepared using different solvents, including chlorobenzene, p-

xylene and 1,2-dichlorobenzene, at various concentrations (5 mg/mL-12 mg/mL). Next,

we used spin-coating, drop casting, and solution-shearing [15, 36-39, 43] methods to

deposit 4T-TMS thin films. For the solution-shearing method depicted in Figure 2.6

(experimental setup placed in a fume hood), briefly, a few drops of semiconductor

20 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

solution were cast onto a pre-heated PTS-modified wafer (2-4 cm2) and covered with a

dewetting OTS-modified top wafer. The top wafer (one quarter of a 5-inch wafer in our

experiments) was then translated by an electric-controlled syringe pump at a constant

rate relative to the bottom device wafer, gradually uncovering the sandwiched solution

which quickly evaporats leaving behind a polycrystalline thin film seeding from the

shearing wafer frontier. Compared to a prior blade coating technique [44], this

deposition method permits controlled solvent evaporation under different deposition

temperature. Moreover, it regulates the deposition rate and growth direction of the

crystalline domains within the film. Solution-shearing processed devices are then placed

on a hot plate for ~10 mins and transferred to a vacuum oven at 80 °C for overnight to

remove the residual solvent. Finally, 40 nm Au source and drain electrodes were

thermally evaporated onto the 4T-TMS films through a shadow mask to complete the

SPOFETs.

Figure 2.6 Illustration of the solution shearing process and analysis of the shearing forces during the film growth.

Based on the above material characterization and the device structure design, we are

able to understand the simplified energy level diagrams for the 4T-TMS SPOFET as

shown in Figure 2.7. Without considering interface dipole effects, Fermi level pinning

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 21

effects [45], and the change of metal work function due to contamination at the

interface of Au and 4T-TMS, we estimate a Schottky barrier height of ~0.2 eV which is

adequate for charge injection from the source to the channel.

(a)

(b)

E

-5.10eV

Vacuum Level

φm φm

4T-TMS

-5.10eV

LUMO: -2.58eV

HOMO: -5.27eV

Au Au

E χ

Ev=-5.17eV

EfEi

Ec=-4.05eV Eg=

8~9eV

0.95eV

SiO2

Vacuum Level

4T-TMS

LUMO: -2.58eV

HOMO: -5.27eV

PTSSi

Figure 2.7 Simplified energy band diagrams without considering interface dipole effect for the top-contact 4T-TMS SPOFET on silicon substrate along (a) source-to-drain channel direction and (b) gate-to-film direction. The HOMO/LUMO levels are based on the measurement as described in Section 2.1.1.

22 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

2.1.3 Comparison of Different Solution Deposition Methods

Spin-coating deposition is widely used for polymer semiconductors due to its simplicity

and capability to produce uniform films [8]. However, this method is not ideal for

fabricating high-quality small-molecule semiconductor films due to small molecule’s

lower solution viscosity and higher crystallinity. As a comparison, we fabricated 4T-TMS

SPOFETs by spin-coating from a chlorobenzene solution at both room temperature

and elevated temperature as shown in Figure 2.8 (a) and (b), respectively. These films are

mostly amorphous and rough, giving a low field-effect mobility typically on the order of

10-3 cm2/Vs. Another method to deposit small-molecule semiconductor film from

solution phase is drop-casting, which is essentially the concept used for the ink-jet

printing of organic electronic devices. Drop-casting normally renders undesirable

randomness in crystalline domain orientation within the film as shown in Figure 2.8 (c)-

(d), consequently sacrificing the performance uniformity.

Notably, Figure 2.8 (e)-(f) show the cross-polarized optical micrographs of typical

4T-TMS thin films as deposited by the solution shearing method under optimized

conditions, where the clear birefringence indicates that well-oriented polycrystalline

domains are successfully formed within the sheared films and they tend to have an

elongated shape along the shearing direction. Tapping mode atomic force microscopy

(AFM) image shown in Figure 2.8 (g) over the good film part further corroborates the

well-oriented crystalline domains.

2.1.4 3-D Statistical Process Optimization

Process optimization is critical in achieving high quality films and high performance

devices. Return to the solution shearing process as shown in Figure 2.6, we can find out

at least two apparently competitive mechanisms in determining the crystalline domain

growth, evaporation and crystallization, both ongoing at the shearing wafer frontier.

Moreover, mechanical shearing force might also play a role in the film growth. Without

knowledge of the crystallization and evaporation kinetics parameters for the involved

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 23

Figure 2.8 Cross-polarized optical micrographs of the 4T-TMS thin films deposited by (a)-(b) spin-coating method, from 5 mg/mL solution in chlorobenzene at (a) room temperature or (b) elevated temperature of ~80ºC; (c)-(d) drop casting method (c) with or (d) without solvent annealing; (e)-(f) solution shearing method with (e) 4T-TMS solution concentration: 6 mg/mL in xylene, deposition temperature: 112 ºC, shearing speed: 0.10 mm/s and (f) 4T-TMS solution concentration: 6 mg/mL in chlorobenzene, deposition temperature: 78.6 ºC, shearing speed: 0.10 mm/s. (g) AFM tapping-mode height image of a local film shown in (e).

24 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

materials and processes thus being difficult for a numerical process modeling at this

moment, we evaluate the primary processing conditions which dominate the growth

output, and optimize them alternatively by systematical experiment design and statistical

data analysis. It can be qualitatively found from the crystallization kinetics theory [46]

and the evaporation kinetics theory [47] that at least the following parameters should be

taken into account to study the shearing process: the solvent type, solution

concentration, shearing speed, and deposition temperature. Therefore, we first prepared

the 4T-TMS SPOFETs under 24 different conditions (noted as “BL-1” here) as listed in

Table 2.1, with 50 devices for each combination of the above four processing variables.

Table 2.1 Experiment design of the first 24 different processing conditions (“BL-1”) toward systematic study and optimization of the shearing process.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 25

The 4T-TMS SPOFET devices were tested using a Keithley 4200-SCS

semiconductor parameter analyzer. For fair comparison with the electrical performance

reported in previous literature, linear curve fitting of DSI vs. VGS in the saturation

regime [refer to Equation (2.1) ] is used to extract significant device parameters as

suggested by the IEEE standard for the characterization of organic transistors [48],

μ

μ μ∂ ∂

≅ + − →∂ ∂

1 1 12 2 2

DS satsat i i GS T sat i

GS GS

I W W WC C V V CV L L V L

(2.1)

where IDS is the drain current, μsat the effective field-effect mobility in the saturation

region, Ci the gate dielectric capacitance per unit area (here Ci≈10 nF/cm2), W/L the

device channel size (here L=13-50 μm, W=960-1000 μm), VGS the applied gate voltage,

and VT the nominal threshold voltage. We calculate the overall μsat and current on/off

ratio (Ion/Ioff) for the SPOFET samples (10%-30%) on each wafer listed in Table 2.1.

High Ion/Ioff with the average value exceeding 106 and a narrow variation is obtained. The

effective mobility, however, depends more heavily on the specific processing condition,

showing an average that spans over an order of magnitude from 0.015 cm2/Vs to 0.17

cm2/Vs for different processing conditions.

A compiled statistical analysis of the 162 SPOFET samples is illustrated in Figure

2.9 and Figure 2.10, which shows the frequency distribution of the average μsat and Ion/Ioff

with respect to various processing conditions. Spline interpolation algorithm is

employed to generate the plots to facilitate comparison. Ideally, optimized conditions

should result in higher frequency in the high-μsat or Ion/Ioff zone. From Figure 2.9 we note

the following:

• For all processing parameters including solvent, solution concentration, shearing

speed and deposition temperature, the first and second highest frequency fall into

the mobility range of 0.01-0.1 cm2/Vs and 0.1-1 cm2/Vs, respectively, with

negligible frequency in the lower mobility zone. This indicates the importance of

process optimization as the frequency in the highest mobility range is obviously

dwarfed due to a large portion of unfavorable processing conditions;

26 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

Figure 2.9 Compiled statistical analysis of 162 SPOFET samples in “BL-1”, which shows the frequency distribution of the average mobility (μsat) with respect to different (a) solvents; (b) solution concentrations; (c) shearing speeds; and (d) deposition temperatures as listed in Table I. Spline interpolation algorithm is employed to generate the plots to facilitate comparison

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 27

Figure 2.10 Compiled statistical analysis of 162 SPOFET samples in “BL-1”, which shows the frequency distribution of the average Ion/Ioff for different (a) solvents; (b) solution concentrations; (c) shearing speeds; and (d) deposition temperatures as listed in Table I. Spline interpolation algorithm is employed to generate the plots to facilitate comparison.

28 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

• Decline rate of the frequency from the range 0.01-0.1 cm2/Vs to 0.1-1 cm2/Vs

reflects the effect of different processing conditions on SPOFET performance.

Superior conditions are expected to show slighter frequency drop as outlined by

the dashed arrow in each plot;

• More devices based on xylene rather than chlorobenzene appeare in the region of

0.1-1 cm2/Vs as shown in Figure 2.9 (a); Nevertheless, the difference between

these two solvents is insignificant;

• Higher solution concentration (12 mg/mL vs. 6 mg/mL) give rise to fewer high-

mobility devices as revealed in Figure 2.9 (b);

• Shearing speed must be controlled properly as shown in Figure 2.9 (c); 0.26 mm/s

in this batch of experiment leads to dramatic drop in the frequency of high-

mobility devices as compared to 0.10 mm/s;

• Appropriate deposition temperature is critical for high-mobility devices as

suggested by Figure 2.9 (d). The optimal temperature is roughly 60% of the

solvent boiling point in centigrade (Tb).

In contrast to the complex frequency distribution of the mobility, Ion/Ioff shows a

more consistent distribution shape irrespective of the processing parameters as shown in

Figure 2.10. Most Ion/Ioff fall in the range of 106-107 with negligible portion in other

ranges. It is not surprising that 4T-TMS SPOFETs show such high on/off ratios

considering that 4T-TMS has a relatively deep HOMO level of -5.27eV. Further

discussions and analysis thus will be only focused on the mobility.

Based on the above analysis, our second batch of 22 different processing conditions

(noted as “BL-2” here) was designed to verify and improve the top 10 high-performing

conditions in the first batch as well as to further optimize the shearing speed and

solution concentration. Specifically, [0.05, 0.10, 0.20] mm/s and [5, 8, 10] mg/mL were

carried out for the shearing speed and solution concentration, respectively. The same

temperatures as in “BL-1” were adopted in “BL-2” since they were found to be an

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 29

adequate range as seen from Figure 2.9 (d). Besides chlorobenzene and xylene, 1,2-

dichlorobenzene was added into the solvent category of “BL-2”.

Figure 2.11 (a) shows similar μsat frequency distribution for the three different

solvents. Combining the results from “BL-1”, we argue that the small variation of the μsat

frequency distributions for different solvents may be within experimental errors. It’s not

conclusive how the solvents systematically affect the μsat based on these observations.

In order to find out the optimal mobility with respect to the input processing

parameters, three-dimensional surface and contour plots of the average μsat as a function

of different shearing speeds, solution concentrations, and deposition temperatures as

shown in Figure 2.11 (b)-(d) are generated using a Renka-Cline interpolation gridding

algorithm based on the data points from 172 measured SPOFETs in “BL-2”. These

plots are consistent with each other as well as consistent with the results from “BL-1”,

and they clearly indicate the optimal processing window for the shearing process: 5.9-8.5

mg/mL for the 4T-TMS solution concentration, 0.09-0.14 mm/s for the shearing speed,

and 55%-70% of solvent boiling point (in centigrade) for the deposition temperature.

Qualitatively, these findings can be interpreted in following:

• Solution concentration effect. It is straightforward that over diluted solution is

difficult to yield oriented crystalline films since crystallization or drying of the film

is slower than the solution shearing [see Figure 2.12 (a) for an example]. On the

other hand, high concentration regularly induces aggregate spots and nonuniform

film on the device substrate as shown in Figure 2.12 (b).

• Shearing speed effect. We consistently observed that nonuniform film with rough

surface appear at the late stage of the deposition if the shearing speed is

excessively low [see Figure 2.12 (c)]. This can be attributed to the faster solvent

evaporation than the film crystallization in the circumstance with a low shearing

speed. On the other side, high shearing speed can give rise to poorly oriented

crystalline domains since crystallization may be slower than the shearing [see

Figure 2.12 (d)].

30 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

Figure 2.11 (a) The frequency distribution of μsat for the three different solvents used in “BL-2”. Spline interpolation algorithm is employed to generate the curves to facilitate comparison. (b)-(d) Three-dimensional surface and contour plots of the average mobility (μsat) as a function of different shearing speeds, solution concentrations, and deposition temperatures. Here we applied Renka-Cline interpolation gridding algorithm and experimental data points of 172 SPOFETs in “BL-2” for the plotting.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 31

Figure 2.12 (a)-(b) Solution concentration effect, (c)-(d) shearing speed effect; and (e)-(f) deposition temperature effect in the solution shearing process, as revealed by the bright-field or cross-polarized microscopy images.

32 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

• Deposition temperature effect. Optimized deposition temperature is essential for

yielding uniform and high-performance SPOFETs. Compared to the optimal

processing window, lower temperature gives rise to little oriented crystalline film

along the shearing direction [Figure 2.12 (e)], being equivalent to high shearing

speed effect or low concentration effect as revealed in Figure 2.12 (d) and (a).

These effects are all associated with the solvent overflow at the shearing wafer

frontier during the deposition. If a high deposition temperature is applied, the

films then tends to be thicker and to become rough as shown in Figure 2.12 (f). In

this case, shadow mask method for the deposition of source/drain electrodes is

generally difficult to get control due to the uneven film surface, which makes the

gap irregular and degrades the SPOFET performance.

Figure 2.13 shows the film thickness as measured by profilometry [Veeco Dektak]

on the wafers listed in Table 2.1, and reinforces the aforementioned findings and

conclusions: higher temperature increases the sheared film thickness, while higher

shearing speed tends to make the film thinner and less uniform. Interestingly, we also

found in this experiment that the film thickness for those high-performance devices (μsat

above ~0.1 cm2/Vs) is typically 20-50 nm.

A1 A2 A3 A4 A5 A6 B1 B2 B3 B4 B5 B6 C1 C2 C3 C4 C5 C6 D1 D2 D3 D4 D5 D60

20

40

60

80

Film

thic

knes

s (n

m)

Wafers in Batch "BL-1" (listed in Table 2.1)

Typical film thickness range of high-performance devices

Figure 2.13 Measured film thickness on the selected wafers of “BL-1” as listed in Table 2.1.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 33

2.1.5 Optimized SPOFET Performance on Silicon Substrate

The previous statistical modeling and data analysis on the systematic experiments has

enabled us to apply the optimal processing conditions to fabricate high-performance,

solution-sheared 4T-TMS SPOFETs. Typical films deposited under the optimized

conditions are exemplified in Figure 2.8 (e)-(f) and briefly discussed in Section 2.1.3. In

this section, we present the electrical characteristics, performance uniformity and

environmental stability of the 4T-TMS SPOFETs fabricated on the silicon substrate

under the optimized solution-shearing conditions.

Figure 2.14 shows the representative transfer and output characteristics of the 4T-

TMS SPOFETs with two different channel lengths (LA=35 μm and LB=16 μm), both of

which were prepared on the same wafer under the following conditions:

• 4T-TMS solution concentration (C) in xylene is 8 mg/mL;

• Shearing deposition temperature (T) is 84 ºC, 60% of the xylene boiling point in

centigrade;

• Solution shearing speed (R) is 0.10 mm/s;

Both devices A and B showed a linear relationship for DSI ~VGS at room

temperature. The IDS~VDS curves represent excellent transistor behavior without

noticeable signs of concave starting induced by large contact resistance or zero-point

shifting due to significant leakage current, even though neither the gate nor the

semiconductor is patterned here. It should be noted that the shorter channel devices

show an inferior subthreshold slope due to a higher leakage current from the channel.

This can be attributed to the contamination from the worse controlled short-channel

shadow mask during the gold deposition process. For devices A and B, the extracted μsat

are ~0.2 cm2/Vs and ~0.3 cm2/Vs, with Ion/Ioff ~106 and ~7×105, and the maximum

transconductances 2.6 μS and 8.1 μS, respectively. According to recent review articles

[49-50], these devices are among the higher performing SPOFETs based on soluble

organic semiconductors, especially on asymmetric small molecules.

34 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

40 0 -40 -8010-10

10-9

10-8

10-7

10-6

10-5

10-4

10-3

VGS (V)

Dra

in C

urre

nt I D

S (A

)

0.000

0.003

0.006

0.009

0.012

0.015Device A: LA=35 μm (W/L=28)

I DS0.

5 (A0.

5 )

Vds=-100V

0 -20 -40 -600

-10

-20

-30

D

rain

Cur

rent

I DS

(μA

)Drain Voltage VDS(V)

VGS=-60V

VGS=-50V

VGS=-40V

VGS=-30V

Device A: LA=35 μm (W/L=28)

40 0 -40 -8010-9

10-8

10-7

10-6

10-5

10-4

10-3

Ci~10nF/cm2

VGS (V)

Dra

in C

urre

nt I D

S (A

)

0.000

0.005

0.010

0.015

0.020

0.025Device B: LB=16 μm (W/L=60)

I DS0.

5 (A0.

5 )

Vds=-100V0 -20 -40 -60

0

-20

-40

-60

-80

Dra

in C

urre

nt I D

S (μ

A)

Drain Voltage VDS(V)

VGS=-60V

VGS=-50V

VGS=-40V

VGS=-30V

Device B: LB=16 μm (W/L=60)

Figure 2.14 Electrical characteristics of solution-shearing processed 4T-TMS SPOFETs with two different channel lengths (LA=35 μm and LB=16 μm), both of which were prepared on the same wafer under the following conditions: 4T-TMS concentration C=8 mg/mL in xylene, deposition temperature T=84 ºC, shearing speed R=0.10 mm/s. (a) Device A shows effective mobility μsat ~0.2 cm2/Vs, Ion/Ioff ~106. (b) Device B shows μsat ~0.3 cm2/Vs, Ion/Ioff ~7×105; the relatively poor subthreshold slope may be caused by contamination from the short-channel shadow mask during the gold deposition.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 35

To understand our SPOFET performance uniformity under a specific processing

condition, we plot in Figure 2.15 the mobility and Ion/Ioff distribution of 18 measured

samples on the same wafer as devices A and B. These devices, with channel lengths of

13-50 μm and channel widths of 960-1000 μm, show an overall μsat=0.20±0.06 cm2/Vs

and an average Ion/Ioff>106.

0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.400

15

30

45

60 Gaussian Fit

Freq

uenc

y (%

)

Effective Mobility µsat (cm2/Vs)

0

20

40

60

80 Gaussian Fit

1010103

Freq

uenc

y (%

)

Ion/Ioff102 105104 109108107106

Figure 2.15 The effective saturation mobility (μsat) and Ion/Ioff distribution of 18 measured samples fabricated on the same wafer as for devices A and B shown in Figure 2.14. The red line is based on the Gaussian fitting model.

36 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

In both Figure 2.15 (a) and (b), the following shifted Gaussian model is invoked to

fit the experimental mobility and current on/off ratio distribution,

2

220 2

cx xAy y e σ

σ π

−−= + (2.2)

where y is the frequency (%), x is the μsat in linear scale or Ion/Ioff in logarithmic scale, y0

and A is a constant, σ is the standard deviation and xc is the mean value. For the above

Gaussian fitting to the distribution of μsat [Figure 2.15 (a)] and Ion/Ioff [Figure 2.15 (b)],

the determination coefficients (R2) are 0.94847 and 0.96949, respectively, suggesting that

both distributions are generally well described by the shifted normal function, albeit they

can still be skewed by a few exceptional high/low-performing devices. Satisfactory

uniformity with σ/xc= 21.31% for the μsat and σ/xc= 7.59% for the Ion/Ioff have been

obtained, suggesting promising applications of these SPOFETs in large area electronics.

If further taking into account the mobility scaling effects due to channel length variation

as discussed in Chapter 4, we can find even higher intrinsic performance uniformity here.

For practical applications, environmental stability is an important issue remained to

be addressed for many high mobility solution processable organic semiconductors

including pentacene and anthradithiophene derivatives [51]. The stability of the 4T-TMS

SPOFETs was tested under ambient air and light conditions. Figure 2.16 shows the

stability measurement results. Up to 180 days after the fabrication, the devices essentially

showed no degradation in mobility or current on/off ratio, indicating good air-stability

for the 4T-TMS. The slight fluctuation of the electrical performance over time was

induced by the relative humidity variation between 37% and 60% when the

measurement was carried out. In order to test light-induced degradation, another wafer

was continuously exposed to a fluorescent lamp (~40 W) since the 23rd day. The

mobility decreased by 53% after 13 days of continuous fluorescent light illumination,

suggesting a photooxidation degradation mechanism for the 4T-TMS SPOFETs. This is

consistent with the measured HOMO energy level of -5.27eV as shown in Figure 2.3.

Overall, compared to pentacene and many high-mobility acene-derivative OFETs which

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 37

start to degrade shortly after exposure to air [51], these 4T-TMS SPOFETs exhibit

favorable environmental stability.

(a)

0 20 40 60 80 100 120 140 160 180 2000

20

40

60

80

100

120

Cur

rent

On/

Off

Rat

io

Time (Day)

Nor

mal

ized

Mob

ility

(%)

Normalized Mobility, Device-1 Normalized Mobility, Device-2 Current On/Off Ratio, Device-1 Current On/Off Ratio, Device-2

103

106

109

(b)

0 5 10 15 20 25 30 35 400

20

40

60

80

100

120

Normalized Mobility Current On/Off

Time (Day)

Nor

mal

ized

Mob

ility

(%)

103

106

109

Cur

rent

On/

Off

Rat

io

Ambient Air & Occasional Light Continuous Exposure to Fluorescent Light (~40 W)

Figure 2.16 (a) Stability measurement of solution-shearing processed 4T-TMS SPOFETs stored in ambient air with occasional exposure to room light. (b) Stability measurement of five 4T-TMS SPOFETs fabricated on the same substrate, stored in ambient air with occasional exposure to room light for 23 days, and later continuously exposed to a fluorescent light (~40 Watts). For both plots, the average mobility measured at the first time is normalized to 100%; the zero time corresponds to the date of device fabrication.

38 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

2.2 Flexible SPOFET Device: Performance and Discussion

In Section 2.1, we have investigated the electronic properties of the organic

semiconductor, 4T-TMS, and optimized the solution-shearing process by systematic

experiment design in conjunction with 3-D statistical data modeling/analysis. By virtue

of comprehensive understanding of the material and the process, we have demonstrated

high performance 4T-TMS SPOFETs on silicon substrate [15]. In this section, we

resume the discussion on our high-performance flexible 4T-TMS SPOFETs fabricated

on plastic substrate as introduced at the beginning of this chapter [see Section 2.1]. The

optimized solution-shearing process for devices on silicon substrate has been

successfully applied to their counterparts on plastic.

For the flexible SPOFETs shown in Figure 2.2, the semiconductor film was

deposited directly on the PBTS-modified flexible PET substrate from a 5 mg/mL 4T-

TMS solution in chlorobenzene at 80 ºC with a shearing speed of 0.10 mm/s, followed

by curing in a vacuum oven at 70 ºC overnight. To overcome the low thermal

conductivity issue for plastic substrates, we heated the plastic in an oven rather than on

hotplate to guarantee the PBTS-modified surface is at the desirable temperature during

the shearing deposition. Again, large crystalline domains with an elongated shape along

the shearing direction and widths up to several hundred micrometers can be clearly

identified from the birefringence observed under a cross-polarized optical microscope

[see Figure 2.17 inset], completely following the results on silicon substrate. To gain a

better understanding of the film microstructure, we performed out-of-plane X-ray

diffraction (XRD) measurements of solution-sheared 4T-TMS thin films which

exhibited distinct diffraction peaks even to the sixth order at 2θ = 4.37º (001), 8.74º

(002), 13.10º (003), and 26.30º (006) as shown in Figure 2.17, indicating a long-range,

highly-ordered π-stacking within the large crystalline domains. The sharp primary peak

at 2θ = 4.37º corresponds to a d spacing of 20.20 Å. Judging from a computed

molecular length of 4T-TMS molecule using MM2 energy minimization and PM3

geometry optimization (20.99 Å), the tilt angle relative to the substrate surface of 74º is

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 39

estimated. This result suggests that solution-sheared 4T-TMS films have highly

favorable molecular orientations for charge carrier transport along the channel [52-53].

0 5 10 15 20 25 300.0

5.0x104

1.0x105

1.5x105

2.0x105

2.5x105

3.0x105C

ount

s/s

2θ (degrees)

(001)

(002)(003) (006)

Au E

lect

rode

Arr

ay Shea

ring

Figure 2.17 Out-of-plane X-ray diffraction (XRD) spectra of solution-sheared 4T-TMS thin film. The inset is a cross-polarized optical microscopy image of the highly crystalline film on plastic PET substrate.

The electrical characteristics of our flexible 4T-TMS SPOFETs were measured and

analyzed in the same way as for the devices on silicon substrate [see Section 2.1.4],

except that here Ci≈22 nF/cm2 and W/L=1000 μm/50 μm. The flexible SPOFETs are

found showing excellent transistor behavior at a bias down to -1.5 V [see Figure 2.18 (a)

and (b) for the transfer and output characteristics, respectively] with a near zero turn-on

voltage and a low VT of 0.5-0.7 V. Quantitative analysis of the turn-on/threshold

voltage is beyond the scope of this work; instead, we propose a simplified energy band

diagram shown in Figure 2.18 (c) which suggests the ideal turn-on/threshold voltage is

attributed to an excellent alignment between the work function of PEDOT:PSS, the

HOMO and Fermi level of 4T-TMS, and the dipole moment of PBTS SAM at the

semiconductor-dielectric interface. The PEDOT:PSS film not only acts as a buffer layer

avoiding detrimental indium diffusion from ITO to the polymeric dielectrics in our

40 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

SPOFETs, but also reduces the threshold voltage of the transistors by ∆Vt=0.4-0.5 V

per its higher work function (5.2-5.3 eV) [54] than that of ITO (4.8 eV).

0.5 0.0 -0.5 -1.0 -1.5-13

-12

-11

-10

-9

-8

-7

sqrt

(I ds) (

A0.5)

Gate Voltage Vgs (V)

Log (

I ds) (

A) S ~80mV/dec

0.0000

0.0001

0.0002

0.0 -0.5 -1.0 -1.5

Vgs

-0.6 V-0.9 VDr

ain C

urre

nt I ds

(nA)

Drain Voltage Vds (V)

-1.2 V

0

-20

-40

0 2 4 6 8 10 120

200

400

600

Operating Voltage (V)

Subth

resh

old S

lope (

mV/de

c)

1.0E-9

1.0E-8

1.0E-7

1.0E-6

1.0E-5

Ion (A)

101102103104105106

on/of

f rati

o

Figure 2.18 (a) Transfer [Vds=-1.5 V] and (b) output characteristics of the flexible 4T-TMS SPOFETs. (c) A simplified energy band diagram clarifying the device physics responsible for the ideal turn-on voltage. (d) Device performance with respect to different operating voltages.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 41

Significantly, at an operating bias of -1.5 V, our best devices exhibit: (1) a field-effect

mobility up to 0.22 cm2/Vs, in good agreement with the highly crystalline film structure

and efficient molecular microstructure for charge transport as discussed above, (2) an

on/off ratio over 5×104, reaffirming the effectiveness of the bilayer PVP-EAD dielectric

structure on control of leakage and off current, and (3) a subthreshold slope of ~80

mV/dec. To our knowledge, this is the lowest subthreshold slope reported for flexible

organic transistors and getting close to the theoretical limit of ln10×kT/q=60 mV/dec

at room temperature. For comparison, Figure 2.19 summarizes the representative

flexible SPOFET performances reported in the past decade. Previously, Klauk et al. [55]

and Cho et al. [29] demonstrated flexible OFETs with state-of-the-art subthreshold

slope down to 100-140 mV/dec, either based on novel SAM dielectrics on oxidized

aluminum/silicon [55] or ion-gel dielectrics [29]. Our work here represents a different

approach based on simple polymeric dielectrics to achieve exceptional subthreshold

characteristics for flexible OFETs. As briefly mentioned at the beginning of this chapter,

the transfer characteristics of organic transistors are largely influenced by both the

interface and bulk traps which lead to degradation of the subthreshold performance [31].

Assuming roughly that the density of deep bulk trap states Nbs and interface trap states

Nis are independent of energy, we can estimate the maximum Nbs and Nis to be ~3×1015

cm-3eV-1 and 6×1010 cm-2eV-1, respectively, by [56]

2

2

1ln10 /

1 /ln10 /

iis

ibs s

S CNkT e e

S CNkT e e

ε

≤ −⋅

≤ −⋅

(2.3)

where e is the elementary charge, k the Boltzmann constant, T the temperature, Ci the

dielectric capacitance per area, and εs the semiconductor dielectric constant (here the

relative constant set to 3 for a simple oligothiophene [57]). Compared to typical one-

order higher interface/bulk trap density reported for OFETs [31], these results indicate

excellent dielectric-semiconductor interface quality for our flexible 4T-TMS SPOFETs

42 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

based on the PBTS-modified PVP-EAD dielectric. Additionally, they also suggest a very

low trap density in the solution-sheared 4T-TMS films.

0.1 1 10

10-2

10-1

100

H. Sirringhaus, et al., Science (2000) J. Rogers, et al., IEEE EDL (2000) J. Rogers, et al., PNAS (2001) G. Gelinck, et al., Nat. Mater. (2004) R. A. Street, et al., Mater. Today (2006) D. Gundlach, et al., Nat. Mater. (2008) K. Myny, et al., ISSCC (2009) M. Bohm, et al., ISSCC (2006) S. Lee, et al., Org. Electron. (2008) H. Yan, et al., Nature (2009) J. H. Cho, et al., Nat. Mater. (2008) H. Klauk, et al., Nature (2007) This work

Mob

ility

(cm

2 /Vs)

(Subthreshold Slope)-1 (dec/V)

Figure 2.19 A review on the representative flexible SPOFET performances (except [54] which is based on vapor process) reported in literature in the past decade. Our work contributes to one of the highest performances, with a record-low subthreshold slope of ~80 mV/dec.

For higher driving current applications, we further investigated the electrical

characteristics at different operating voltages (i.e. initial Vds for Ids-Vgs test). It is not

surprising that the subthreshold slope increases slightly and the on/off ratio degrades

with the operating voltage as shown in Figure 2.18, primarily due to the use of a

common gate structure [see Figure 2.2 (a)]. A relatively large overlap capacitor between

the drain/source and gate contributes leakage to the off current which increases

exponentially with the applied electric field, while the on current changes approximately

according to a square law. We anticipate an improved operating-bias-dependence of the

subthreshold characteristics by patterning the gate and further optimization of the

polymer dielectric properties.

CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR 43

2.3 Summary

We have shown that flexible SPOFETs on rough plastic substrates with a charge carrier

mobility up to 0.22 cm2/Vs, a turn-on voltage of near 0 V, and a record-low

subthreshold slope of ~80 mV/dec are possible in ambient conditions. These

exceptional characteristics are attributed to: (1) a novel device stacking architecture with

a conducting polymeric gate electrode and a double layered dielectric composed of low-

temperature cross-linked poly(4-vinylphenol) (PVP); (2) a low interface trap density

achieved by modifying the PVP dielectric surface with a phenyl-terminated self-

assembled monolayer from 4-phenylbutyltrichlorosilane (PBTS); and (3) controlled

crystallization of a small-molecule organic semiconductor 4T-TMS film with favorable

charge transport microstructure and a low bulk trap density as deposited by a statistically

optimized solution-shearing process.

In addition, we present the systematic experiment design and the corresponding 3-D

statistical modeling/analysis which provides a general guideline for process optimization

for fabricating high-performance SPOFETs.

Future work in this part includes analytical or quantitative modeling for the shearing

process based on evaporation kinetics and crystallization kinetics, enhancing the device

tolerance to higher driving-current operation through gate patterning, and integrating

such SPOFETs in flexible electronic circuits and systems.

44 CHAPTER 2. SOLUTION-PROCESSED FLEXIBLE ORGANIC TRANSISTOR

45

Chapter 3

Interface Engineering by Self-

Assembled Monolayer

It doesn't matter how beautiful your theory is, it doesn't matter how smart you are. If it

doesn't agree with experiment, it's wrong.

- Richard Feynman

As briefly mentioned in Chapter 2, organic transistor performance is closely associated

with the surface condition of the dielectrics, or, equivalently, the interface property

between the semiconductor and the dielectrics. This chapter focuses on studying the

interface effects and relevant interface engineering by self-assembled monolayers (SAM).

3.1 Background and Motivation

Looking back to history, even since the first transistor invented by John Bardeen, Walter

Brattain, and William Shockley at Bell Labs in 1947, people have been realizing the

crucial role of the surface states in determining the field-effect transistor behavior [58].

46 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

Numerous scientists and engineers have practically spent their whole life in eliminating

the defects to improve the semiconductor-dielectric interface for traditional MOSFET.

Indeed, this interface is equally important for organic thin-film field-effect transistor,

simply based on the fact that most of the charge carriers induced by the gate voltage in

the field-effect transistor, regardless of inorganic or organic, are accumulating in a very

thin channel close to the semiconductor-dielectric interface, as illustrated in Figure 3.1.

Any defects, trap states, or imperfections are anticipated to influence charge transport

along the interface.

Figure 3.1 Schematic illustration of the gate voltage induced charge carrier layer, which is close to the semiconductor-dielectric interface.

A widely adopted measure to modify the dielectric surface for organic transistors is

based on self-assembled monolayers (SAM) as first proposed for vapor processed

pentacene transistor [41, 59]. The basic concept is illustrated in Figure 3.2: organosilane

molecules with a silane headgroup react with the hydroxyl-terminated dielectric surface,

spontaneously assemble and polymerize in coverage of the entire dielectric surface to

remove originally unfavorable trapping sites or defects. It has been shown that [15, 36,

41-42, 59-65], by adequate employment of SAMs with desired molecular structure, one

is able to modify the interface states distribution, charge carrier density, surface energy,

and film growth mode and crystallinity, all of which may contribute to improved device

performance and reliability.

As earlier representative attempts to understanding the SAM effects on the organic

transistor, Salleo et al. [65] investigated solution-processed poly-9,9’ dioctyl-fluorene-co-

bithiophene (F8T2) thin-film transistors with the dielectric surface modified by different

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 47

SAMs, and suggested that SAM induced mobility enhancement may be involved with

molecular interactions between the polymer and the SAM. This work also pointed out

the possibility of using dipolar SAM molecule to control the threshold voltage for the

transistor. Later, Kobayashi et al. [64] and Pernstich et al. [63] did a more systematic

study on the SAM-dipole effects on the charge carrier density and the threshold voltage.

These works provide pioneering insights into general physical and chemical effects

collectively arising from the SAM molecules. However, the SAMs being used in each

particular study possess different functional groups which are in direct interaction with

the organic semiconductor. Depending on the chemical properties, it is not impossible

for the functional groups to distort the apparently observed dipole moment relevant

physical effects. Furthermore, the SAM quality as deposited even from the same process

can vary considerably for the SAMs with different functional groups. This issue used to

be ignored and raises further complications in interpreting the experimental

observations on the organic transistor’s electrical characteristics. Other works [60-62]

have also explored the SAM effects based on the molecules with the same functional

end group, typically –CH3, but with different intra alkyl-chain lengths. These studies

exclude the chemistry effects of the functional groups from the physical effects,

however, no dipole relevant physical effects were separately investigated. It is thus still

unclear how the specific functional group’s chemistry effect and the dipole relevant

physical effect may each contribute to the interface engineering for organic transistors.

SiO

OO

Si

OO

Si

OO

Si

OO

CFF

F

Figure 3.2 Schematic illustration of the concept of dielectric surface modification by self-assembled monolayers.

48 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

In this work, through careful selection of a group of phenyl-terminated SAMs, we

manage to decouple the dipole moment relevant physical effect from the surface

chemistry effect of the functional groups, to elucidate how the performance and

reliability of organic transistors are controlled by interfacial SAMs [66-67].

The chemical structures of the phenyl-SAM molecules used for this study are shown

in Figure 3.3. The major distinctions of this study are highlighted below:

PTS PETS PBTS PHTS PAPTS

Si O

OO

NH

Si ClCl

ClSi Cl

Cl

Cl Si

Cl

Cl Cl SiClCl

Cl

Figure 3.3 A group of phenyl-terminated SAM molecules used for the interface engineering study in this work. The intra alkyl chain length increases from PTS, PETS, PBTS, to PHTS, PAPTS is comparable to PBTS except that an –NH- group replaces a –CH2- in the intra part of PBTS. (PTS: phenyltrichlorosilane, PETS: phenethyltrichlorosilane, PBTS: 4-phenylbutyltrichlorosilane, PHTS: 6-phenylhexyltrichlorosilane, PAPTS: N-phenylaminopropyltricholorosilane)

• As introduced in Chapter 2 [see Figure 2.5], phenyl-terminated SAMs are

particularly favorable for SPOFETs and can be superior to other SAMs including

the methyl-terminated OTS which is widely used for vapor processed organic

transistors. Therefore, comprehensive understanding of the phenyl-terminated

SAM effects will enable us to further improve the SPOFET performance;

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 49

• With the same functional aromatic end group and the consistent intra alkyl chain

from PTS to PHTS, we are able to exclude the interference of the SAM functional

group’s chemistry effects to the SAM dipole physical effects. This differs from

previous studies and will enable us to decouple these two effects for individual

identification;

• With the only difference of an intra-chain atom between PBTS and PAPTS, we

are able to testify the chemical effects per specific SAM molecular structures and

provide direct evidence of SAM chemistry effects on the SAM quality and the

relevant transistor performance and reliability.

• Ambient atmosphere including H2O and O2 has been shown to have strong

influences on the transistor performance [68], bringing additional artifact sources

in the observed SAM effects. Therefore, we pay particular attention in the device

fabrication and measurement, all in N2 atmosphere or vacuum without exposure

to air during the processing, thus excluding any possible intervence of the ambient

environment in analyzing the SAM effects.

• Consistent quality of the phenyl-SAMs being investigated in this work is ensured

by a dedicated new deposition process, thus eliminating the artificial effects

resulted from different interfacial SAM quality.

3.2 Phenyl-SAM Processing and Characterization

3.2.1 Spin-coating Process and Mechanism

In general, organosilanes can be either vapor deposited or solution deposited onto the

substrate surface to render the SAM structure. Vapor process regularly yields a smoother

[6] but low-density and amorphous SAM layer [69]. For the latter case, a typical protocol

is immersing the substrate in a solution of organosilanes and solvents for at least a few

hours. This method, though being easier in processing, is prone to rendering a rough

50 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

multilayer structure due to the high susceptibility of silanes to the presence of water [6].

In reality, the silane soaking process also suffers from contamination most likely from

the backside of the wafer. Additionally, for flexible electronics applications, immersing

the flexible substrate in the silane solution faces more challenges.

Recently, Nie et al. proposed a simple spin coating process using nonpolar solvents

such as trichloroethylene (TCE) and chloroform with a dielectric constant of ~4 to

deliver full-coverage octadecylphosphonic acid (OPA) SAM on oxide surfaces [70].

Later, Ito et al. extended this approach to deposit crystalline ultrasmooth alkylsilane

SAM such as OTS on silicon oxide surface, giving rise to very high mobility for vapor

processed pentacene and C60 transistors [71]. The same mechanism involved in both

processes is that a raft of hydrophilic headgroups of OPA and OTS in the nonpolar

solvent like TCE spontaneously aggregate and align on the liquid drop surface,

providing an ideal environment to fast react with the hydroxyl-terminated oxide surface

during the spin coating process [70]. However, both OPA and OTS are methyl-

terminated long alkyl chain molecules and their based SAMs are more favorable in vapor

processed organic transistors instead of in SPOFETs as discussed in the background

section. In this work, we extend the method by Nie et al. [70] and Ito et al. [71] to

develop a simple spin coating process for the deposition of high-quality phenyl-

terminated SAMs which play particularly important roles in SPOFETs [72].

The basic process and its relevant mechanism are illustrated in Figure 3.4. We

emphasize that the conditions including the solvent, concentration and spin coating

rate/time are critical to the success of the SAM deposition, as revealed by the

preliminary film screening from microscopy images shown in Figure 3.5. Of particular

importance is the application of anhydrous toluene as the solvent for the phenyl-

terminated silanes here. Although TCE, chloroform and toluene are all nonpolar

solvents, the former two proven to be ideal for spin-coating of OPA and OTS SAMs

[70-71] are found being inappropriate for phenyl-terminated silanes. This observation

indicates the importance of the aromatic group of toluene in interacting with the

functional endgroup of the phenyl-silanes and thus further facilitating the

orientation/alignment of the trichlorosilane headgroup on the liquid surface.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 51

Figure 3.4 Schematic illustration of the spin-coating process for depositing phenyl-terminated SAMs from the solution of phenyl-trichlorosilane in anhydrous toluene. The trichlorosilane headgroup spontaneously aggregate and align on the liquid drop surface, providing an ideal environment for the fast reaction of the trichlorosilane with the hydroxyl group on the substrate surface [70-71]. The anhydrous toluene solvent being selected here is critical to the process. The conjugated aromatic group of toluene is supposed to interact with the functional endgroup of the phenyl-silane and facilitate the alignment of the phenyl-silane on the liquid drop surface.

Figure 3.5 A preliminary screening of different film morphology as deposited by different spin-coating processes for PBTS. Images were taken under bright-field microscope. The solvent, concentration and spin rate/time are all critical to high quality phenyl-SAM.

52 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

The aligned tricholorsilane headgroup on the liquid surface reacts fast with the

hydroxyl group on the substrate surface and the excessive solution is removed during

the spin process. Subsequent post-processing by treating the spin-coated substrates in

hydrochloric acid vapor for overnight is necessary to allow further hydrolysis and

siloxane polymerization on the surface. The optimized process details are given in the

Appendix section in this dissertation.

3.2.2 Spin-coated Phenyl-SAM Characterization

The phenyl-terminated SAMs as deposited by the same optimized spin-coating process

are characterized by techniques including water contact angle (CA), atomic force

microscopy (AFM), ellipsometry, and grazing incidence X-ray diffraction (GIXD).

Figure 3.6 shows the measured water contact angle [Edmund Scientific goniometer] on

each phenyl-SAM modified SiO2 surface. For PTS, PETS, PBTS, and PHTS with similar

molecular structure, the water contact angle increases monotonically with the intra-alkyl-

chain length from ~70º to ~90º, indicating relatively low surface energy and high density

of the SAMs as deposited by the spin coating process. For PAPTS which contains a -

NH- in the intra-chain, a lower contact angle is observed over the comparable PBTS

where only -CH2- appears in the chain. This effect may arise from the SAM morphology

difference as induced by the molecular structures during the spin coating process, and

the stronger dipolar interaction between the PAPTS SAM and the polar water molecule.

As shown in Figure 3.7, AFM measurement [Veeco Nanoscope] reveals that the

phenyl-terminated SAMs as deposited by the optimized spin-coating process are

generally showing remarkably high degree of smoothness, with typical roughness of 0.1-

0.2 nm (rms) for PTS/PETS/PBTS/PHTS and 0.2-0.3 nm (rms) for PAPTS. No

appreciable difference among the morphology of PTS/PETS/PBTS/PHTS SAMs is

observed. However, a relatively rougher film surface is found for PAPTS, indicating the

SAM molecular structure with specific chemical functional groups in affecting the SAM

quality even based on the same process. This underscores the importance of selecting

appropriate SAM structures in studying the SAM effects for organic electronic devices.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 53

PTS PETS PBTS PHTS PAPTS50

60

70

80

90

100

Mol

ecul

ar L

engt

h, b

y M

OPA

C9

(Å)

Con

tact

Ang

le (0 )

Si

Cl

Cl Cl Si

Cl

ClCl

SiO

O

O

NH

SiCl

Cl

Cl

SiCl

Cl

Cl

Exception for N-included chain

0

2

4

6

8

10

12

14

16

Figure 3.6 Measured water contact angle on phenyl-SAM modified SiO2 surfaces. The molecular length of each SAM molecule is computed by PM3/MOPAC9 model and shown for comparison purpose here.

Si

Cl

Cl Cl

Si O

OO

NH

Figure 3.7 Representative phenyl-SAM film morphology for (a) PTS/PETS/PBTS/PHTS and (b) PAPTS, as deposited by the optimized spin-coating process. For PTS-series SAM, the typical roughness is 0.1-0.2 nm (rms); for PAPTS SAM, the roughness is 0.2-0.3 nm (rms). This indicates the spin coating process yields excellent phenyl-SAM quality.

54 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

Long alkylsilanes such as OTS (~2 nm) have been reported to form a well packed,

crystalline SAM on the amorphous SiO2 surface under appropriate processing

conditions [71]. However, this phenomenon is not observed for the relatively shorter

phenyl-silane based SAMs as investigated in this work. GIXD experiments performed at

Stanford Synchrotron Radiation Lightsource (SSRL) on these phenyl-SAM films

deposited on native SiO2 do not show crystalline features in the diffraction pattern [see

Figure 3.8], indicating the phenyl-SAMs are amorphous or disordered, probably due to

the weaker van der Waals interactions between neighboring molecules for the short

phenyl-SAM molecules (<1.4 nm).

SiCl

Cl

Cl

Figure 3.8 GIXD image of the PHTS SAM on native silicon oxide surface as deposited by the spin coating process. By courtesy: Dr. Eric Verploegen.

The phenyl-SAM layer thickness is measured using ellipsometry [Sopra Bois-

Columbes] and shown in Figure 3.9. In addition, we compute the SAM molecule length

with PM3/MOPAC9 model and the Gaussian Package 03’, B3LYP/6-31G(d) model.

Both simulations yield similar results on the molecular length. Therefore, it is possible to

estimate the average tilt angle for the SAM molecule on the surface by

cos /d Lθ = (3.1)

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 55

where θ is the tilt angle to the surface normal direction, d is the measured film thickness,

and L is the molecular length, for which the Gaussian simulation is used here. As shown

in Figure 3.9, the average tilt angle for PTS/PETS/PBTS/PHTS is ~40º-55º.

Interestingly, the PAPTS is found giving a considerably larger SAM thickness and

negligible tilt angle compared to the phenyl-SAMs without the nitrogen atom.

PTS PETS PAPTS PBTS PHTS --0

5

10

15

20

25

30

35

40

-100

-75

-50

-25

0

25

50

75

100

Mol

ecul

ar L

engt

h or

SA

M T

hick

ness

( Å) Measured SAM Thickness, d

Simulated Molecule Length, L (PM3/MOPAC9) Simulated Molecule Length, L (Gaussian)

Tilt

Angl

e, θ

(0 )

cos (θ)= d/L

Figure 3.9 Measured phenyl-SAM thickness (d) using ellipsometry and the simulated molecule length (L) based on PM3/MOPAC9 model or Gaussian Package 03’, B3LYP/6-31G(d) model. The average title angle of the SAM molecule with respect to surface normal direction, θ, is estimated by cos (θ)=d/L, where L is based on the Gaussian simulation here.

3.3 Phenyl-SAM Effects in Organic Transistors

3.3.1 PBTTT Transistor Structure and Device Fabrication

To systematically investigate the SAM effects on the organic transistor performance and

reliability, we fabricated poly(2,5-bis(3-alkylthiophen-2-yl)thieno[3,2-b]thiophene)

56 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

(PBTTT) transistors with the structure shown in Figure 3.10, which incorporate

different phenyl-SAMs introduced above at the dielectrics-semiconductor interface.

PBTTT is a high-mobility, solution-processable p-type material, exhibiting liquid-

crystalline phase under thermal annealing conditions [73]. PBTTT film is spin-coated

onto the phenyl-SAM modified 300 nm SiO2 following the procedures reported before

[73], rendering relatively high uniformity for its transistor performance. All of the device

fabrication and measurement were performed in N2 or vacuum environment without

exposure to ambient O2 and H2O, thus excluding the environment induced artifacts [68]

in the observed SAM effects.

Thermally grown SiO2

N++ Si

4T-TMS Active LayerD S

PBTTTD S

PTS PETS PBTS PHTS PAPTS

Si O

OO

NH

Si ClCl

ClSi Cl

Cl

Cl Si

Cl

Cl Cl SiClCl

Cl

PBTTT

Figure 3.10 Device structure of PBTTT transistor incorporating different phenyl-SAMs at the dielectric-semiconductor interface. Courtesy: PBTTT was provided by Dr. Iain McCulloch and Dr. Martin Heeney (formerly in Merck).

3.3.2 Device Electrical Characteristics

In general, PBTTT transistors based on the spin-coated phenyl-SAMs exhibit excellent

transistor behavior with satisfactory transfer and output characteristics [see Figure 3.12

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 57

for an example]. Relatively high mobility (>0.4 cm2/Vs) and on/off ratio (>105) are

obtained for PBTS and PHTS based devices.

20 0 -20 -40 -60 -8010-9

10-8

10-7

10-6

10-5

10-4

Gate Voltage VGS (V)

Dra

in C

urre

nt I D

S (A

)

0.000

0.003

0.006

0.009

0.012

0.015

I DS0.

5 (A0.

5 )

VDS=-80V

0 -20 -40 -60 -800

-20

-40

-60

-80

-100

-120

-140

D

rain

Cur

rent

I DS

(μA

)

Drain Voltage VDS (V)

VGS=-80V

VGS=-60V

VGS=-40V

VGS=-20V

Figure 3.11 Representative (a) transfer and (b) output curves for the PBTTT transistors based on the spin-coated phenyl-SAMs. The results shown here are specifically based on PBTS.

Figure 3.12 shows the overall average mobility and threshold voltage for the PBTTT

transistors incorporated with different phenyl-SAMs or without any SAM at the

dielectric-semiconductor interface. It is found that,

• For PTS/PETS/PBTS/PHTS with alkyl intra chain only, they all improve the

transistor performance compared to that without SAM modification on the

dielectric surface. In this series of phenyl-SAMs, longer intra chain gives rise to

higher mobility, in consistence with the water contact angle and surface energy

measurement results as shown in Figure 3.6;

• PAPTS, despite having comparable molecular structure to the other phenyl-SAMs

where only a –CH2- is replaced by –NH-, yields significantly lower mobility. This

phenomenon clearly indicates the chemical effects resulted from the specific SAM

structure, and reiterates the field-effect mobility is not simply determined by the

surface energy, otherwise the mobility for PAPTS should be similar to that for

PTS as judged from Figure 3.6;

58 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

• The threshold voltages of the PBTTT transistors with phenyl-SAMs are in

opposite region to that without SAM, and they are apparently dependent on the

particular SAM structure.

PAPTS SiO2 PTS PETS PBTS PHTS0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

Mob

ility

(cm

2 /Vs)

no-SAM

PAPTS SiO2 PTS PETS PBTS PHTS-15

-10

-5

0

5

10

15

20

Thre

shol

d Vo

ltage

(V)

no SAM

20 0 -20 -40 -60 -8010-9

10-8

10-7

10-6

10-5

10-4

VGS (V)

Dra

in C

urre

nt I D

S (A

)

0.000

0.003

0.006

0.009

0.012

0.015

I DS0.

5 (A

0.5 )

Vds=-80V

Figure 3.12 (a) Mobility and (b) threshold voltage of PBTTT transistors incorporated with different phenyl-terminated SAMs at the dielectric-semiconductor interface.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 59

It should be emphasized that hysteresis measurement provides ample information

on the interface property. The general mechanisms and the corresponding effect on the

hysteresis of the transfer curve are illustrated in Figure 3.13.

VGS (V)

Dra

in C

urre

nt I D

S (A

)

0(1)

VGS (V)

Dra

in C

urre

nt I D

S (A

)

0(2)(3)(4)

Figure 3.13 General mechanisms responsible for the hysteresis effect in the organic transistor. Only (1) the charge injection from semiconductor to the interface induces counterclockwise hysteresis in the transfer curve. The other mechanisms including (2) charge injection from the gate electrode, (3) residual dipole caused by slow polarization in the bulk dielectric, and (4) mobile ions in the dielectric contribute to the counterclockwise hysteresis in the transfer curve.

The measurement results on the hysteresis of our devices with different phenyl-

SAMs are shown in Figure 3.14. For all the devices, we only observed counterclockwise

hysteresis depicted in Figure 3.13 (note: the reversed VGS axis in Figure 3.14), which

indicates that charge injection from the semiconductor to the interface dominates in

these devices. PTS/PETS/PBTS/PHTS are found rendering similar negligible hysteresis

effect in the transfer curve for PBTTT transistors. Moderately large hysteresis can be

seen in devices without SAM modification. However, substantially larger hysteresis is

60 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

found for the PAPTS modified PBTTT transistors. We thus argue that the surface

modification by spin-coated PTS/PETS/PBTS/PHTS yields improved, high-quality

interface conditions, while spin-coated PAPTS SAM tends to give rise to more traps and

defects at the interface. This argument is further supported by bias stress measurement

as shown in Figure 3.15. Considerably larger bias stress effect (threshold voltage shift) is

observed for PAPTS modified PBTTT transistors, suggesting more traps and defects

exist at their dielectric surface, while no appreciable difference on the bias stress effects

for the PTS/PETS/PBTS/PHTS is observed. Therefore, the spin-coated PTS series

SAMs are believed to exhibit similar interface quality.

20 0 -20 -40 -60 -8010-10

10-9

10-8

10-7

10-6

10-5

10-4

VGS (V)

Dra

in C

urre

nt I D

S (A

)

SiCl

Cl

Cl

SAM: PBTS

Vds=-80V

20 0 -20 -40 -60 -8010-11

10-10

10-9

10-8

10-7

10-6

10-5

SiO

O

O

NH

VGS (V)

Dra

in C

urre

nt I D

S (A

)

SAM: PAPTS

Vds=-80V

60 40 20 0 -20 -40 -60 -8010-10

10-9

10-8

10-7

10-6

10-5

10-4

VGS (V)

Dra

in C

urre

nt I D

S (A

)

No SAM

Vds=-80V

Figure 3.14 General hysteresis effect for PBTTT transistors modified with different phenyl-SAMs. PTS/PETS/PBTS/PHTS show no appreciable difference in the hysteresis, therefore only a representative I-V hysteresis curve is given here as adopted from PBTS based transistors.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 61

1 10 100 1000 100000.01

0.1

1

10

100

Thre

shol

d Vo

ltage

Shi

ft, -d

Vth

(V)

Stress Tiime (s)

PTS PETS PBTS SiO2 PAPTS PHTS

VGS=-60 VTrecovery=200 s

Si O

OO

NH

Figure 3.15 Bias stress effect on the PBTTT transistors with different phenyl-SAMs. Considerably larger bias stress effect is observed for the PAPTS modified interface. The applied gate bias for the bias stress measurement is VGS=-60V. For each I-V measurement during the bias stress test, a recovery time of 200 seconds is given to allow the fast reversible traps recovered.

3.3.3 Threshold Voltage Control by SAM Dipole Moment

The threshold voltage of organic transistors is apparently related to the specific SAM

molecular structure. To understand how the threshold voltage is controlled by the SAM,

we show in Figure 3.16 the energy band diagram of the PBTTT transistors with SAM

modification layer at the dielectric-semiconductor interface. The dipolar SAM creates an

additional electric field at the interface, changes the accumulation/depletion regime

when the gate voltage is zero, or equivalently, bends the energy alignment, and thus

tunes the threshold voltage. A simple model to resolve this effect is shown in Figure

3.17. The induced surface charge density (QS) by the addition of the SAM dipole at the

dielectric interface is described as

62 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

Figure 3.16 Simplified energy band diagram of the PBTTT transistor incorporating the SAM modification layer at the dielectric-semiconductor interface. The SAM dipole creates an additional electric field normal to the interface, bending the energy alignment and thus tuning the threshold voltage.

Figure 3.17 A simple model to describe the dipole moment effect in tuning the surface charge density and transistor’s threshold voltage. The SAM dipole (p) on the dielectric surface induces additional surface charge (QS). Θ is the tilt angle of the dipole moment.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 63

cos SAMs

SAM

p NQL

⋅ Θ⋅= (3.2)

where p is the SAM dipole moment, Θ is the tile angle of the dipole relative to the

surface normal, NSAM is the SAM density, and LSAM is the SAM molecular length. The

corresponding shift of the threshold voltage (ΔVt) is found to be

0

Tox

ox

ox

ox

xdxe Qs

TVtC

ρ+

Δ =−

∫i

(3.3)

where ρox is the mobile charge density in the dielectric, Tox is the dielectric thickness, Cox

is the dielectric capacitance per area, e is the elementary charge.

We emphasize that the threshold voltage can be only compared in this model for the

devices with similar interface SAM quality, otherwise the distinct trapping effects would

affect the observed threshold voltage and make the comparison unfair. Therefore, based

on our observations and discussions on the PAPTS of which the chemical effects

significantly influence its SAM property and distinguish it from the other alkyl-chain

only phenyl-SAMs, we exclude this molecule in the following analysis.

The dipole moment for each molecule (p0), SAM molecule density (NSAM), and SAM

molecular length (LSAM) are simulated using Gaussian package and PM3/MOPAC9

model as listed in Table 3.1, and the tilt angle Θ is estimated to be 30º-60º based on the

SAM characterization shown in Figure 3.9 as well as the dipole simulation. One should

note that the dipole moment in the monolayer form (p) is attenuated over the isolated

molecule due to Coulomb interaction of parallel dipoles [74]. Therefore, an attenuation

factor of 5 is assumed here [74], i.e. p=p0/5.

64 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

Table 3.1 SAM molecule dipole, molecule length, and SAM density as simulated using Gaussian package and PM3/MOPAC9 model.

SAM

molecule

Molecule dipole

(p0) [debye]

Molecule length

(L) [Å]

SAM density (NSAM)

[Å2/molecule]

PTS 1.31 7.01 19.46

PETS 1.67 8.93 25.32

PBTS 1.69 11.41 21.80

PHTS 1.67 14.48 26.68

Figure 3.18 shows the measurement results and the simulation results for the

threshold voltage shift with respect to the SAM dipole and SAM induced surface charge

density. In general, experimental results on both p-type PBTTT transistors and n-type

amorphous phenyl-C61-butyric acid methyl ester (PCBM) transistors show that the

threshold voltage shift is linear to the SAM dipole-induced surface charge density, in

excellent agreement with the theoretical model shown in equation (3.3).

However, discrepancy between the experiment fitting line and the theoretical line for

the slope is still noticeable. We emphasize that an important mechanism has not been

considered yet to date, i.e. strong external interfacial electric field as induced by the gate

voltage would distort the inherent dipole moment of the SAM molecule as illustrated in

Figure 3.19, thus influence the model fitting in (3.3). This effect, namely electric

shielding tensors effect, must be given careful consideration in future studies. From the

experiment point of view, in-situ and real-time measurement of the SAM dipole on an

oxide surface under the applied external electric field would give direct evidence of the

electric shielding tensors effect. In the meantime, it is also possible to model this effect

using Ab-initio simulation in the future.

CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER 65

0 1x1012 2x1012 3x1012 4x101230

25

20

15

10

5

0

-5

-10

-15

0 1x1012 2x1012 3x1012 4x101230

25

20

15

10

5

0

-5

-10

-15

0 1x1012 2x1012 3x1012 4x101230

25

20

15

10

5

0

-5

-10

-15

PTSPETS

PBTSPHTS

Theoretical Line (Slope=-e/Cox)

Experimental Fitting Line

PTS

PETSPBTS

Experimental Fitting Line

Thre

shol

d Vo

ltage

(V)

Surface Charge Density (cm-2)

Theoretical Line (Slope=-e/Cox)

PHTS

PBTTT

PCBM

Figure 3.18 Measurement results vs modeling results for the threshold voltage shift with respect to the surface SAM dipole and SAM induced surface charge density. Both p-type PBTTT transistors and n-type PCBM transistors are measured in this work.

66 CHAPTER 3. INTERFACE ENGINEERING BY SELF-ASSEMBLED MONOLAYER

Figure 3.19 Schematic illustration of the proposed electric shielding tensors effect of the SAM at the dielectric-semiconductor interface: strong external interfacial electric field as induced by the gate voltage would distort the inherent dipole moment of the SAM molecule, thus influencing the performance and reliability of the organic transistors.

3.4 Summary

We have systematically investigated dipole moment related physical effects and

chemistry effects of the SAM at the organic semiconductor-dielectric interface. Through

careful selection of a group of phenyl-terminated SAMs, we elucidate how the

performance and reliability of organic transistors are controlled by interfacial SAMs. In

addition, we briefly introduce a spin-coating process for depositing high-quality phenyl-

terminated SAMs for organic electronics applications.

67

Chapter 4

Device Physics and Universal

Modeling of Organic Transistor

In physics, your solution should convince a reasonable person. In math, you have to convince a

person who's trying to make trouble. Ultimately, in physics, you're hoping to convince Nature.

And I've found Nature to be pretty reasonable.

- Frank Wilczek

So far we have presented the experimental demonstration of high-performance flexible

organic thin-film transistors based on solution process and systematically investigated

the electrical performance and reliability controlled by the semiconductor-dielectric

interfacial monolayer. In this chapter, we’ll study the device physics of organic

transistors with focus on charge injection effects at the metal-organic interface and

charge transport properties in the organic semiconductor film. A universal device model

capturing many of the physical effects observed for organic transistors will be

introduced [75].

68 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

4.1 Background and Motivation

As an important figure of merit for transistors, the transition frequency, defined by

fT ∝ μFETV/2πL2, reflects circuit switching speed/bandwidth and strongly depends on

the extrinsic field-effect mobility (μFET)1 as well as the channel length (L). Therefore,

appropriate downscaling of the organic thin-film transistor, resembling that of

traditional silicon MOSFET which has been substantially benefited from the

sophisticated device scaling theories [1], is also crucial for the development of organic

electronic circuits and systems. Earlier works have explored organic transistors with

channel lengths from hundreds of micrometers down to sub-100 nm [76-84].

Experimental results show that the extrinsic field-effect mobility of these transistors is

channel-length variant, typically, with a lower value for sub-micrometer-channel devices.

As a consequence, the advantages originated from dimensional scaling to organic

circuits are severely compromised. Even for organic transistors with channel lengths

above micrometers, it is still not uncommon that the mobility decreases with shortening

of the channel length [81, 85-90]. Such channel-length dependence of the extrinsic

mobility has been suggested qualitatively to contact resistance effects; nevertheless,

analytical and quantitative studies remain in lack.

On the other hand, an inverse trend to the above mobility scaling behavior has been

reported for organic transistors more recently [15, 27, 43, 86, 91-92]. As briefly

mentioned in Chapter 2, for instance, we found an extrinsic field-effect mobility

decreasing with channel length for solution-processed, top-contact, polycrystalline

small-molecule organic TFTs [15, 43]. Hamadani et al. [91] and Verlaak et al. [92]

observed an higher effective mobility in the saturation regime for shorter channel

polythiophene-/pentacene-based TFTs and they attributed this to the electric field

dependence of charge transport. Gundlach et al. [27] and Smits et al. [86] found similar

mobility scaling behavior for contact-controlled polycrystalline acene-based TFTs and

quinquethiophene-based self-assembled-monolayer field-effect transistors (SAMFET),

1 In literature, it is also termed “apparent” or “effective” mobility, the value of which is extracted from the transfer curve in linear or saturation regime

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 69

respectively. However, the origins are interpreted differently. For Gundlach et al.’s

devices with contact-induced polycrystalline active layer, decreased disorder and reduced

density of grain boundaries in the shorter channel device is believed in lowering the

activation energy for charge transport and accordingly enhancing the intrinsic mobility

[27]. While for Smits et al.’s SAMFET work, they argued that the film coverage and two-

dimensional percolation are responsible for the mobility scaling characteristics [86].

The apparent paradox arising from a variety of literature reports demands that the

mobility scaling behavior of organic transistors should be resolved in a more complete

way, both analytically and quantitatively. Also, fundamental understanding of the elusive

mobility scaling behavior is essential for technological improvement of the device

performance and crucial for scaled organic circuits design.

In this work, we propose and develop a universal physical model for organic

transistors toward resolving the mobility scaling behavior, clearly elucidating their

underlying physical mechanisms by taking into account both contact effects at the

metal-organic interface and charge carrier transport properties in the active organic

semiconductor layer. We introduce an extended polaronic mobility-edge (EPME) model

which includes the electric field dependence of the charge carrier mobility, namely,

Poole-Frenkel (PF) like exponential dependence of the carrier mobility on the electric

field. Contact effects are described in a system of diffusion-limited charge injection into

a disordered organic film with Gaussian density of states (DOS) at the metal-organic

semiconductor interface. This universal model then combines both space charge limited

conduction (SCLC) and gate voltage induced charge accumulation to describe the

transistor operation. It successfully enables us to account in a single formula for most of

the significant physical phenomena observed so far for organic semiconductors. Further

analysis reveals that, depending on the physical origins including carrier injection barrier

height, localized trap states energy distribution, electric field dependence of carrier

mobility, and channel length dependence of film crystallinity, the mobility scaling curve

with respect to the channel length ranging from a few nanometers to hundreds of

micrometers can be monotonically upwards, or slowly get saturated, or, surprisingly,

exhibit an overshoot region.

70 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

We stress that, beyond resolving the mobility scaling issue, our device model is also

versatile in explaining many other elusive physical phenomena for organic thin-film

transistors, such as the contact resistance effect and the surface potential profiles along

the channel which have been experimentally probed but yet poorly understood. Since

there is no commercially available device simulator dedicated for organic transistors

(unlike well-developed industry standard device simulator, e.g. Sentaurus, for traditional

silicon MOSFET), we implement the universal model into a 1-D device simulator for

organic TFTs. The simulation results of various electrical characteristics are in excellent

agreement with experimental observations.

4.2 Universal Modeling of Organic Transistors

There are a variety of structures available for standard organic TFTs [6], which can be

broadly categorized as bottom-contact and top-contact devices shown in Figure 1.2. A

compact model to describe the electrical characteristics of organic TFTs seemingly relies

on their particular structure [93-94], yet, the underlying physical mechanisms dominating

the device operation are consistent, and, are possible to describe in a generalized 1-D

manner as depicted in Figure 4.1.

Figure 4.1 A generalized 1-D representation of the physical process dominating the device operation for organic transistors.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 71

Charge carriers are first injected from the source electrode to the organic film

through the metal-organic semiconductor interface, then cross a transition region in the

vicinity of the source electrode before reaching the active channel layer. Eventually, the

charge carriers are extracted at the drain electrode through the metal-organic interface

immediately after traversing the other transition zone close to the drain. In bottom-

contact devices, the transition regions arise from enhanced structural disorder of the

organic semiconductor film deposited along the electrode edge as revealed by atomic

force microscopy (AFM) [95], and they tend to give rise to multifold effects, including:

• Charged traps enriched in the depletion zone screen out the gate voltage [96-97]

and lead to considerably lower free carrier concentration than that in the channel;

• Enhanced disorder and lower carrier concentration in the amorphous or

microcrystalline transition region reduce the effective carrier mobility;

• Disordered film structure induces inhomogeneities and energy level fluctuations at

the metal-organic semiconductor interface [97-104].

These effects must be carefully considered and taken into account for the device

modeling. Note that similar transition regions and relevant effects can be also identified

in top-contact devices as confirmed by both experimental observations [105] and

simulation results [93, 106-107]. The transition width (d(tr)), however, differs from in top-

contact and bottom-contact structures since the former is more heavily related to the

organic film thickness due to the fact that most charge carriers need to cross the entire

film to reach the interfacial channel. While for top-contact devices, metal evaporation

and organic film surface contamination are anticipated to cause conjugation break and

traps creation [105-106] in the neighborhood of the contacts, thus contribute to the

transition width.

Based on the above analysis, our universal device model incorporates contact effects

[88, 90, 93-97, 108-115] at the metal-organic interface as well as in its neighboring

transition region, and charge transport effects from the transition region to the active

channel. Since the contact resistance is normally found less pronounced at the drain for

charge extraction [112, 114], for simplicity, we only consider the lumped contact effects

72 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

at the source where charge injection occurs. For consistency and generality, we focus on

p-type organic transistors throughout discussions in this chapter. The results apply to n-

type devices by the same token.

4.2.1 Charge Injection at the Metal-Organic Interface

Although charge injection at the metal-organic semiconductor interface has been

investigated extensively with diversified theories [116], finding an appropriate model

suitable for organic transistors is not an easy task. One of the most classical models to

treat injection from metal to semiconductor or insulator is based on Richardson-

Schottky (RS) thermionic emission theory [117], in which the current density (J) is

expressed as

3*

022 3

/4exp exp

2b

RSe FeemJ kT

kT kTπεφ

π= ⋅ − ⋅ (4.1)

where e is the elementary charge, m* the carrier effective mass, ħ the reduced Planck

constant, k the Boltzmann constant, T the temperature, φb the real barrier height

between the metal and the semiconductor or insulator, which can deviate from the ideal

Schottky barrier due to Fermi-level pinning effect [45] as discussed in Chapter 5, F0 the

electric field at the interface, and ε the dielectric constant of the semiconductor or

insulator material. The second exponential term in equation (4.1) represents the barrier

lowering by ∆φb=(eF0/4πε)1/2 at a distance of x0=(e/16πεF0)1/2 away from the interface

due to external electric field and the image force as illustrated in Figure 4.2. The RS

model reflects the maximum charge injection current and doesn’t take into account

carrier diffusion or inelastic scattering induced backflow current inside low-mobility

materials, including most organic semiconductors as discussed in this work.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 73

Metal Fermi level

Metal-Organic Interface

b∆ b

External electric field:-eF0x

HOMO0 x

Image force: -e2/16πεx

Gaussian localized States Mean electrostatic potentialU(x)=e B-eF0x-e2/16πεx

x0

Interface dipoles:Fermi-level pinning

Figure 4.2 The energy level diagram of the metal-organic semiconductor interface.

We therefore invoke diffusion-limited thermionic emission model for charge

injection at the metal-organic interface [116, 118-122]. In this framework, charge

injection has been either treated by considering an extra drift-diffusion process [118-

120], or, alternatively, in conjunction with surface recombination in an image potential

[116, 121-122]. Indeed, the model can be described in the following general form,

3

00 0

/4' , exp expbtr

e FeJ eN F TkT kT

πεφμ β= ⋅ ⋅ − ⋅ (4.2)

with N0 being the effective density of states (DOS) for the organic semiconductor, μ’(tr)

the carrier mobility in the transition region and close to the contact interface where

electric field is equal to F0. Note that the mobility here is dependent on the electric field,

temperature, carrier concentration as well as the film microstructure [52-53, 123], and

differs from that in the active channel region. We’ll come back to this point in next

section. β(F0, T) is a function in terms of the electric field and temperature, and its exact

74 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

formula depends on the particular derivation approach. A more accurate and unified

description at both low [118-119] and high [119] electric field can be deduced from

Scott et al.’s injection-recombination model [121-122] as follows:

2 2 2 220

0 0 3 30 0 0

0 0

3 1/40 0

4 4 4, 4 1

,

2,

eFk T kT k T eF T Fe F e eF e F kT

F T F

kTF T eFe

πε πε πεβπε

β

β π επ

⎧⎪⎪⎪⎪⎪⎪⎪ = ⋅ + − +

=⎨

=

Simmons'/Emtage & O'Dwyer's Low-Field Limit [100-101]:

Emtage & O'Dwyer's High-Field Limit [101]:

⎫⎪⎪⎪⎪⎪⎪⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎬⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎪ ⎪⎩ ⎭

(4.3)

In order to justify equation (4.3) for modeling our metal-organic semiconductor

system, it is required to satisify

0 02 3* 2

' ,2 1tr

RS

N F TJJ m kT

μ βπ

⋅= ≤

⋅ (4.4)

to ensure diffusion being a contact dominating factor. Figure 4.3 shows the calculated

boundary condition for diffusion-limited charge injection based on equations (4.3) and

(4.4), where, for organic semiconductors at room temperature, N0 and m* is assigned

typical value of 1021 cm-3 [124] and 3×me0 [125], respectively. For an electric field of 1-

100 MV/m, which is the typical field strength at the source/drain of bottom-contact

organic transistors as estimated from scanning Kelvin probe potentiometry [112, 114],

the boundary mobility at T=300 K is found to be 10-60 cm2/Vs. Clearly, for most

organic semiconductor films [6, 53] or even organic single crystals [126-127], their

carrier mobility readily satisfies this boundary condition.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 75

105

106

107

108

109

10-2

10-1

100

101

102

103

Electric Field, F0 (V/m)

Car

rier M

obili

ty (c

m2 /V

s)

Recombination ModelLow-Field LimitHigh-Field Limit

Figure 4.3 The calculated boundary condition for diffusion-limited charge injection at metal-organic semiconductor interface.

So far, we have quantitatively justified the applicability of diffusion-limited charge

injection theory to model organic transistors from the perspective of pristine metal-

organic interface. However, the disordered nature of organic semiconductor films close

to the contact interface has not yet been considered. As pointed out in literature before

[97, 109, 112, 128], neglect of the disorder feature raises contradictory argument on the

activation energy of the injection current (eφJ). Experimental results showed that eφJ is

lower than the activation energy of carrier mobility (eφμ) [109, 112, 128], apparently in

disagreement with prediction from equation (4.2), where φJ=φμ+φb-∆φb≥φμ satisfies. A

well accepted description to the disordered nature of organic semiconductor films is

based on Gaussian density of localized states [97-104, 116] as illustrated in Figure 4.2. In

this circumstance, the injection current density is found to be

76 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

30

0 0

2

2

320

0 0

/4' , exp exp

1 exp2 / 2

/4/2' , exp exp

tr

b

btr

e FeJ eN F TkT kT

e e de

e Fe kTeN F TkT kT

πεφμ β

φ φ φπσ σ

πεφ σμ β

+∞

−∞

⎡ ⎤⎢ ⎥= ⋅ ⋅ − ⋅ ⋅⎢ ⎥⎢ ⎥⎣ ⎦

⎡ ⎤− −⎢ ⎥⋅ ⎢ ⎥⎣ ⎦

−= ⋅ ⋅ − ⋅

(4.5)

where σ is the Gaussian width of the localized states energy distribution. After taking

into account the Gaussian distribution of the localized states at the interface, we find the

activation energy of injection current is lowered by ∆σ=σ2/2kT≈50-200 meV for

T=300 K, σ=50-100 meV [97, 100-104, 109], thus being consistent with experimental

observations.

We emphasize that it is reasonable to treat charge injection at the metal-organic

interface separately from charge transport in the transition region beyond the interface

[97, 119]. Space charge effect is ignored in dealing with charge injection due to the fact

that the rapid variation of the electrostatic potential takes place within a short distance

from the electrode, typically x0=(e/16πεF0)1/2<10 Å, and the electric field F0 may be

regarded as being constant at the interface [119]. For charge transport in the transition

region with a typical width of d(tr)=20-200 nm [97, 109, 112], space charge effect

becomes pronounced since the field strength (F) in this region no longer satisfies the

condition of F>>2ned(tr)/ε~107 V/m, where n is the carrier concentration [119].

Therefore, we conclude that space charge limited conduction (SCLC) dominates the

charge transport in the organic transition region while diffusion limited charge injection

dominates at the metal-organic interface. As discussed in the following sections, they

naturally converge in the device operation through a self-consistent interfacial electric

field, F0. It has been suggested by Koehler et al. [97] recently that the commonly

observed yet poorly understood contact voltage/resistance for organic transistors may

originate from a combining effect of charge injection at the interface and SCLC in the

transition region. However, the feasibility of either mechanism for organic transistors is

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 77

not justified and the mobility difference in different regions is not considered in their

work.

We now proceed to discuss and model the charge transport process in the active

organic semiconductor layer, including that in the transition region close to the contact

and the major channel region, as shown in Figure 4.4 (a).

Figure 4.4 (a) Schematic illustration of the transition region close to the contact and the channel region in the organic semiconductor film; (b) The DOS profiles of the transition region and the major channel based on the mobility edge model; (c) An equivalent representation of (b) with the critical energy for both regions set to be identical to simplify the mathematical derivation.

78 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

4.2.2 Charge Transport in the Active Layer: EPME Model

Here we discuss charge transport in the active semiconductor layer which is obviously

not a pure organic single crystal. Instead, it is an amorphous or polycrystalline organic

film coupled with disorders, traps and defects. These nonidealities give rise to a broad

distribution of Anderson localized states [129], and, accordingly, band only transport

description is unrealistic.

In general, there are two families of classical models available in dealing with charge

transport process in the disordered small-molecule or polymer films. One is based on

the hopping theory, including Miller-Abrahams’ fixed range hopping model [130] and

Mott’s variable range hopping (VRH) model [131], which calculate thermally activated

tunneling events between localized states. In this theoretical framework, carriers may

either hop to a nearest site with large activation energy or a distant site with low

activation energy. Vissenberg and Matters [132] have successfully applied this concept in

conjunction with a percolation criterion for the film conduction to the carrier mobility

model for amorphous organic TFTs. The other one is based on the mobility edge (ME)

concept [131, 133-134] as illustrated in Figure 4.4, or, equivalently, multiple trapping and

release (MTR) concept [135-136], which defines a critical energy (E0, the “mobility

edge”) separating immobile localized states (μ=0) with a DOS tail in the forbidden

bandgap from the extended band of mobile states (μ=μ0). Trapped carriers become

mobile by thermal excitation into the delocalized band and thereby contribute to the

conduction. This model has been successfully applied to study semicrystalline polymer

transistors by Salleo et al. [133] and later refined by Chang et al. [137] by incorporating

the polaronic effect.

For the work presented here, we choose the ME theoretical framework since

hopping transport theory such as the VRH model doesn’t take into account the nature

of supramolecular ordering and mesoscale microstructure in the semi- or poly-crystalline

organic semiconductor films. Also, ME model has been shown from studies of the

microstructure effect and polaronic effect on the carrier mobility that it provides an

improved fitting to the experimental data over VRH based model [133, 137]. It should

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 79

be noted that, however, previous ME models adopted for organic transistors haven’t yet

considered the electric field effect on the mobility. Since the electric filed dependence of

the mobility becomes pronounced at low temperature or high field, which is the

situation typically found in the transition region as discussed in last section, we therefore

first develop an extended polaronic mobility edge (EPME) model here by taking into

account the field dependence of the carrier mobility. This EPME model for charge

transport is then integrated into our universal device model for describing the organic

transistor operation.

As shown in Figure 4.4 (b), the DOS profiles of the organic semiconductor films are

represented by a delocalization band with an exponential tail of donor-like localized

states decaying into the bandgap [133]. The exponential tail here is an approximation of

the Gaussian profile generally accepted for disordered organic semiconductors. First

principle calculations on the electronic structure of polycrystalline polymers [138] have

explicitly shown that the carrier delocalization in ordered lamellas increases the

bandwidths for both the highest occupied molecular orbital (HOMO) and the lowest

unoccupied molecular orbital (LUMO), consequently, giving rise to a bandgap reduction.

Therefore, considering the different disorder levels in the transition region and in the

channel region as discussed at the beginning of Section 4.2, we propose that the critical

energy defining the mobility edge (E0) is shifted by ∆E0=E0(Ch)-E0(Tr), roughly 0.2-0.4 eV,

between these two regions. The DOS profiles shown in Figure 4.4 (b) are described as

0 0

0 0

1 sgnexp2

1 sgn12

tail

tail tail

tail

tail c

N E E E EDOS EE E

N E E E EE E

− − + −= ⋅ ⋅Δ Δ

− + −+ ⋅ − ⋅Δ

(4.6)

where sgn(x) is the sign function, Ec is a parameter to tailor the DOS profile for the band,

Ntail is the concentration of tail states, ∆Etail=Etail-E0 is the tail width of the localized

states. The DOS for the delocalization band is simply represented with a form for 3-D

free hole gas. The above parameters accompanied by a footnote, “tr” or “ch”, in Figure

80 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

4.4 (b), denote those for the transition region or the channel region, respectively. One

should note that the tail states concentration and band mobility (μ=μ0) are assumed

identical for both regions since the major distinction induced by structural modifications

is the localized tail width [133].

When a gate voltage (VGS) applies, charges tend to accumulate along the organic

semiconductor-dielectric interface. The total charge density, including that of immobile

localized trap states and mobile free carriers, can be approximated as

, i GS ontot GS

C V VN V Th−= (4.7)

where Ci is the dielectric capacitance per area, Von is the turn-on voltage, or, equivalently,

the flat band threshold voltage when charge start to accumulate, which ideally should be

around 0 V [36], h is the first-order approximation of the charge layer thickness (~1 nm)

at the dielectric-semiconductor interface.

On the other hand, Figure 4.4 indicates that the total charge density, Ntot(VGS, T),

and the mobile carrier concentration, Nmob(VGS, T), are represented as follows:

, 1tot GSN V T DOS E F E dE+∞

−∞= ⋅ −∫ (4.8)

0

, 1E

mob GSN V T DOS E F E dE−∞

= ⋅ −∫ (4.9)

where

/1

1 fE E kTF Ee −=+

(4.10)

is the Fermi-Dirac distribution. Combine equations (4.6)-(4.10), we can readily find the

solution to Fermi energy, Ef., and the results apply to both the transition region (“tr”)

and the channel region (“ch”). Figure 4.5 shows the calculated relationship between the

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 81

tail width of localized states (∆Etail or σ), the relative Fermi energy level (Ef-E0), and the

applied gate voltage (VGS) for a typical organic transistor at room temperate. It is noted

that the Fermi level position moves to deeper bandgap with the increased localization

tail width, in agreement with the observation that the effective mobility decreases with

the broadening of the localized states. Also, the applied gate voltage systematically tunes

the Fermi energy level position for every single tail width. Higher gate voltage induces

more free charge carriers in the organic semiconductor films, and, equivalently, shifts

the Fermi energy level closer to the mobility edge. Another important feature as revealed

in Figure 4.5 is that, for the transition region where the tail width of localized states is

larger than that the channel region, e.g. ∆Etail(tr)=100 meV and ∆Etail(tr)=30 meV, its

relative Fermi level tends to be 0.2-0.5 eV deeper. Consider the difference in the critical

energy by ∆E0=E0(Ch)-E0(Tr),~0.2-0.4 eV as mentioned before, the absolute Fermi energy

positions in both regions thus align consistently.

0.0 0.2 0.4 0.6 0.8 1.00.0

0.1

0.2

0.3

0.4

ΔEtail(ch)

ΔEtail(tr)-1 V

-10 V-20 V

-40 V-60 V-80 V

Tail

Wid

th (e

V)

Relative Fermi Position, Ef-E0 (eV)

VGS=-100 V

0.0

0.1

0.2

0.3

0.4

Figure 4.5 The relationship between the tail width of localized states (∆Etail), the relative Fermi energy level (Ef-E0), and the applied gate voltage (VGS). The following conditions are employed for the calculation here: 300 nm SiO2 gate dielectric, Von=0 V, h=1 nm, Ntail=1021/cm3, Ec=40 meV, T=300 K.

82 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

Without incorporating polaronic effect, the overall effective carrier mobility, μeff, is

simply defined as μeff=μ0×(Nmob/Ntot) [133]. In reality, strong electron-phonon coupling

existing in the conjugated organic semiconductor molecules or polymers leads to

polaron formation and thus contributes to the temperature activation of the carrier

mobility. A refined description of the effective mobility has been suggested [137] as

0,, exp,

mob GS aeff GS

tot GS

N V T EV TN V T kT

μ μ= ⋅ ⋅ − (4.11)

where the polaronic effect is represented by an extra exponential term, based on the

assumption that the mobility beyond the mobility edge is thermally activated with a

polaron activation energy of Ea.

Another important factor in determining the carrier mobility behavior is the electric

field dependence, which has been an active research subject for disordered organic

materials. It is now well recognized that this dependence strongly relies on the

temperature and field range as well as the film’s structural properties, such as energetic

or positional disorder levels [139-146]. A routine finding on organic semiconductors,

from both Monte Carlo simulations [139-142] and experiments [143-146], is that within

an intermediate field range, typically 104-106 V/cm, the field-dependent mobility exhibits

the Poole-Frenkel like behavior and can be described in a general form as

1exp n FT

μ γ∝ ⋅ (4.12)

where γ(1/Tn) is an empirical parameter relying on temperature (T), n=1-2 [139, 141-

146]. Note that the lateral field strengths for the transition region and the channel region

shown in Figure 4.4 (a) can be considerably different. Therefore, a rough estimation of

the average field by F=VDS/L (L is the channel length) as frequently seen in literature is

not necessarily correct. We estimate the field strengths here from the scanning probe

potentiometry profiles [112, 114] and find that the field dependence as described in

(4.12) is appropriate in both regions. However, it should be stressed that, due to their

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 83

appreciable difference in the film microstructure and thus energetic/spatial disorder,

γ(1/Tn) for the transition region is expected larger than the less disordered channel

region: γ(tr)(1/Tn)>γ(ch)(1/Tn). This can be inferred from Bässler’s formula [139],

2 20

1n C

T kTσγ = ⋅ −∑ (4.13)

where σ is the energetic width of the Gaussian DOS, ∑ represents the off-diagonal

(spatial) disorder. For disordered organic materials, γ(1/Tn) is normally 0.001-0.005

cm1/2V-1/2 at room temperature. With the increasingly ordered structure (or smaller σ)

and higher temperature, γ(1/Tn) eventually vanishes or even turns negative [140, 143].

The extended polaronic mobility edge (EPME) model proposed in this work

incorporates the effective mobility derived in (4.11) as the prefactor for the field-

dependent mobility described in (4.12). The carrier concentration, temperature, and

electric field-dependent effective mobility is expressed collectively as below:

0, 1, , exp exp,

mob GS aeff GS n

tot GS

N V T EV T F FN V T kT T

μ μ γ= ⋅ ⋅ − ⋅ ⋅ (4.14)

Equation (4.14) is a general form adequate for both the transition region and the

channel region. Here we assume the EPME effective mobility derived for charge carriers

located at the dielectric-semiconductor interface applies to the entire film thickness. This

approximation is judged from the nature of mobility edge model which inherently

relates the film microstructure to the effective mobility. For clarity, we reiterated the

mobility in both regions as follows:

0

, 1, , exp exp,

mob tr GS atr GS tr n

tot GS

N V T EV T F FN V T kT T

μ μ γ= ⋅ ⋅ − ⋅ ⋅ (4.15)

84 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

0

, 1, , exp exp,

mob ch GS ach GS ch n

tot GS

N V T EV T F FN V T kT T

μ μ γ= ⋅ ⋅ − ⋅ ⋅ (4.16)

We emphasize that, in circumstances such as in contact-induced polycrystalline

organic thin film transistors [27], solution-shearing processed polycrystalline organic

transistors [15, 38, 43], etc., the overall film crystallinity is correlated to the channel

length. The channel length dependent crystallinity will thus affect the μ(ch) in (4.16) by an

additional function of channel length. We’ll return to this point when discussing the

mobility scaling behavior in Section 4.3.3.

4.2.3 Unified Model for Transistor Operations

Based on the discussion and analysis in Sections 4.2.1 and 4.2.2, we are now able to

combine all of the effects to describe the transistor operation. Since SCLC dominates in

some local regions, it is necessary to include in the overall drain current (IDS) both the

SCLC current arising from the drain-source voltage (VDS) and the charge accumulation

current arising from the gate-source voltage (VGS). Accordingly, we have

DS SCLC tr accumulation tr SCLC ch accumulation chI I I I I= + = + (4.17)

For SCLC component with negligible diffusion current, it follows Gauss’s law

eF ρε

∇⋅ = (4.18)

and the current transport equation

, , , , xSCLC SCLC SCLC x GS x x GS x

dFI J A WHe F V T F WHF V T Fdx

ρ μ μ ε= ⋅ ≅ = ⋅ ⋅ ⋅ (4.19)

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 85

where ρ is the charge carrier density, x denotes the channel direction from source (x=0)

to drain (x=L) as shown in Figure 4.4 (a), ASCLC is the effective cross sectional area for

the SCLC current. In a first-order approximation, ASCLC =W×H, where W is the

transistor’s channel width, H is the organic semiconductor film thickness.

For charge accumulation induced by the gate voltage at the dielectric-semiconductor

interface, it contributes to the drain current at position x by

, ,accumulation GS x x i GS on xI V T F F C W V V Vμ= ⋅ ⋅ ⋅ ⋅ − − (4.20)

where Vx is the electrical potential at position x resulted from the drain-source voltage,

and it relates to the electric field (Fx) by

x xF V=−∇⋅ (4.21)

and

0 0

L L

x x DSF dx V dx V= −∇⋅ =−∫ ∫ (4.22)

Plug equations (4.15), (4.16), and (4.19)-(4.22) into equation (4.17), we have

0

'0

, 1exp exp,

'

mob GS aDS x xn

tot GS

xx

i GS on x

N V T EI WF FN V T kT T

dFH C V V F dxdx

μ γ

ε

= ⋅ ⋅ − ⋅ ⋅

× ⋅ ⋅ + ⋅ − +∫

(4.23)

Integration of equation (4.23) from x=0 to x=L on both sides yields the drain

current equation for the transistor as follows:

86 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

0 0

'0

,1exp exp,

'

La mob GS

DS x xntot GS

xx

i GS on x

E N V TWI F FL kT T N V T

dFH C V V F dx dxdx

μ γ

ε

⎧⎪⎪= − ⋅ ⋅ ⋅⎨⎪⎪⎩

⋅ ⋅ ⋅ + ⋅ − +

(4.24)

Alternatively, equation (4.24) can be formulized with the transition region and the

channel region described by separate terms,

0 0

'0

0

,1exp exp,

'

1exp exp

trd mob tr GSaDS x tr xn

tot GS

xx

i GS on x

ax ch

N V TEWI F FL kT T N V T

dFH C V V F dx dxdx

EW FL kT T

μ γ

ε

μ γ

⎧⎪⎪= − ⋅ ⋅ ⋅⎨⎪⎪⎩

⋅ ⋅ ⋅ + ⋅ − +

+ − ⋅ ⋅

'0

,,

'

tr

L mob ch GSxnd tot GS

xx

i GS on x

N V TF

N V T

dFH C V V F dx dxdx

ε

⎧⎪⎪ ⋅⎨⎪⎪⎩

⋅ ⋅ ⋅ + ⋅ − +

(4.25)

where d(tr) is the width of the transition region, typically 20-200 nm.

Equation (4.23)-(4.25) are the derived current equation for organic transistors based

on our universal model incorporating both the charge injection effects and charge

transport properties. The model is essentially physics-oriented since each variable

involved has real and particular physical meanings. This model successfully enables us to

account in a single formula for most of the significant physical phenomena observed so

far for organic semiconductor films and their based transistors.

In order to solve the above differential equations to fully simulate the transistor’s

behavior, it is necessary to obtain the boundary conditions. Therefore, we invoke the

injection current density at the source contact, JSD, as given by equation (4.5) in Section

4.2.1. Substitute (4.15) for the mobility μ(tr) in (4.5), we obtain:

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 87

3

0 0 0

2

0

/41exp exp

' /2, exp

aSD tr n

mob tr b

tot tr

eEJ eN FkT T kT

N e kTF TN kT

πεμ γ

φ σβ

= − ⋅ + ⋅

−⋅ ⋅ ⋅ −

(4.26)

Note that we distinguish μ’(tr) in equation (4.5) and μ(tr) in (4.15) by assuming that the

carrier mobility for charge injection at the metal-organic interface, i.e. μ’(tr) in (4.5), is

correlated to the film disorder or localized states tail width by N’mob(tr)/Ntot(tr)∝exp(-σ/kT),

and, is independent of the applied gate voltage. While for the charge transport mobility

in the film, i.e. μ(tr) in (4.15), it is apparently gate voltage dependent since Nmob(tr)/Ntot(tr) is

defined by equations (4.7)-(4.10).

The charge injection current equation (4.26) provides the following boundary

conditions for the transistor’s ordinary differential current equation (4.23),

DS SD inj SDI J A J WHα= ⋅ = ⋅ (4.27)

0 ' 00' 0

x

x xV F dx == − =∫ (4.28)

0 0x xF F == (4.29)

where Ainj is the effective injection area, α is a prefactor determined by the transistor

geometry. In addition, to resolve the full transistor behavior, another similar boundary

condition for F(x=d(tr)) and V(x= d(tr)) should also be employed at the transition and

channel region connection position, x=d(tr).

So far, we have analytically derived the universal device model by incorporating both

charge injection effects and charge transport properties for organic thin-film transistors.

This model has been implemented in C/Mathematica programs to render a full 1-D

88 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

device simulator with most of the important physical effects represented by particular

physics-oriented parameters or variables. In next section, we will discuss the simulation

results based on this model and compare them to our or previously reported

experimental results, and hence to testify the theories for the device modeling and

resolve a few elusive physical phenomena for organic transistors.

4.3 Simulation vs. Experimental Results

For the simulation purpose here, we apply typical values as identified from experiments

for the representative parameters in the model (see Table 4.1). The definitions and the

value settings of these parameters have been discussed and justified in Section 4.2.

Table 4.1 Representative parameters, corresponding physical meaning and typical value used for the simulation based on the universal device model in this work

Parameter Physical Meaning Value

W Transistor channel width 1000 µm

L Transistor length Specified in simulation

H Organic semiconductor film thickness 50 nm

d(tr) Transition region width 100 nm

h Charge layer thickness 1 nm

µ0 Intrinsic mobility 1 cm2/Vs

Ci Gate dielectric capacitance 11.5 nF/cm2 (300 nm SiO2)

T Temperature 300 K

Von Onset voltage 0 V

VGS,VDS Gate-source voltage, Drain-source voltage Specified in simulation

Ntail Concentration of tail states 1021 /cm3

Ec DOS profile tailoring parameter 40 meV

Ea Polaron activation energy 15 meV

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 89

ΔEtail(tr) Localized trap states tail width, in transition ~σ, 50-100 meV

ΔEtail(ch) Localized trap states tail width, in channel 30 meV

γ(1/Tn) Prefactor of mobility’s PF field dependence Specified in simulation

ε Organic semiconductor dielectric constant 3 ε0

φb Metal-organic injection barrier height Specified in simulation

σ Gaussian width of film’s localized states ~ ΔEtail(tr), 50-100 meV

4.3.1 Transfer and Output Characteristics

The basic electrical characterization of organic transistors includes the transfer and

output curves as shown frequently in the previous chapters. Based on the universal

device model introduced in Section 4.2, we simulate the output [Figure 4.6 (a)] and

transfer [Figure 4.6 (b)] characteristics for a typical organic transistor with the parametric

values listed in Table 4.1. It can be found that the device model excellently reproduces

the typical IDS-VGS and IDS-VDS curves for organic transistors operating in both the linear

region and the saturation region.

Particularly, with a relatively large injection barrier height at the metal-organic

semiconductor interface (φb=0.7 eV) and localized trap states distribution (σ=100 meV),

a superlinear, concave starting region is clearly reflected in the output characteristics as

shown in Figure 4.6 (a). This phenomenon has been routinely observed in experiments

for organic transistors with poor contacts or significant contact resistance [95, 109, 147].

To gain a better fundamental understanding on the contact-induced peculiarity, we also

simulate the same device with a smaller injection barrier height of φb=0.2 eV. The

superlinearity in the output curves, though does not entirely disappear, is considerably

alleviated per the injection barrier height reduction as seen from Figure 4.6 (a) inset. In

addition, the driving current is enhanced by ~20% over the device with an injection

barrier height of 0.7 eV. This clearly shows the importance of controlling the metal-

organic interface for low injection barrier height to optimize the device performance.

90 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

The residual superlinearity for even low injection barrier height can be attributed to the

other factor, localized trap states distribution width (σ), as discussed in next section.

Compare the above simulation results to our 4T-TMS SPOFET performance as

presented in Chapter 2, we reaffirm the 4T-TMS SPOFET possesses favorable contacts

at the source/drain with a low contact injection barrier height and a narrow localized

trap states distribution in the organic film.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 91

0 -20 -40 -60 -80 -100

0

50

100

150

VGS=-20 V

VGS=-60 V

VGS=-80 V

VGS=-100 V

I DS (μ

A)

VDS (V)

φb=0.7 eV, σ=0.1 eV

φb=0.2 eV, σ=0.1 eV

VGS=-100 V

0 -10 -20

0

10

20

30

40

I DS (μ

A)

VDS

(V)

(a)

0 -20 -40 -60 -80 -100

0

20

40

60

80

100

120

140

VGS (V)

I DS (μ

A)

10-13

10-11

10-9

10-7

10-5

(b)

I DS (A

)

VDS=-100 V

Figure 4.6 Simulation results of the (a) output and (b) transfer characteristics, based on the universal device model introduced in Section 4.2. In addition to parameters given in Table 4.1, the followings are included: L=10 µm, γ(tr)(1/Tn)= 0.001 cm1/2V-1/2, γ(ch)(1/Tn)=0, (a) φb=0.7 or 0.2 eV, σ=0.1 eV, and (b) φb=0.7 eV, σ=0.1 eV. Note: for the transfer curve simulation, an Ioff

(IDS@VGS=0)=1 pA is assumed per practical situations.

92 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

4.3.2 Surface Potential Profile and Contact Resistance

Surface potential profile along the dielectric-semiconductor interface has been well

studied and understood in traditional silicon MOSFETs [117]. It was until recently that

this topic was recognized to be critical in studying organic thin-film transistors. In fact,

the surface potential profile contains ample information on the transistor operation

status and the contact effects. A few early works have probed this potential profile for

organic transistors using high-resolution conductive probe atomic force microscopy

(AFM) [114], electrostatic force microscopy (EFM) [148], or scanning Kelvin probe

microscopy (SKPM) [112, 149], thus providing a foundation for justifying the electric

field distribution and estimating the contact resistance. However, the fundamental

physics and mechanisms responsible for these experimental observations are still poorly

understood.

Based on our universal model which takes into account both the charge injection

effects at the metal-organic interface and charge transport properties in the active film

layer, we are able to simulate the surface potential profile for the organic thin-film

transistors [see Figure 4.7 (a)]. Devices operating from linear regime (-VGS>-(VDS+VTH))

to saturation regime ((-VGS<-(VDS+VTH))) are explicitly revealed by the surface potential

profile change upon increase of the applied drain voltage. For –VDS>80 V, the channel

area close to the drain is gradually getting to depletion mode for saturation operation.

The channel regions which are in accumulation mode show a seemingly quasi-linear

potential profile and only consume little voltage drop along the channel, in distinct

contrast to the transition regions where a significant voltage drop occurs in the case of

φb=0.7 eV and σ=100 meV. This dramatic voltage consumption in a short distance close

to the source, i.e. the transition region as depicted in Figure 4.4 (a), contributes to the

extrinsically large contact resistance as regularly found in devices with poor contacts. For

direct comparison, Figure 4.7 (b) invokes a representative surface potential profile of the

P3HT-polymer transistor as recorded in experiments by noncontact potentiometry using

scanning Kelvin probe microscope by Bürgi et al. [112]. It can be found that the

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 93

universal model developed in this work excellently reproduces surface potential profiles

as previously probed by experiments.

0 2 4 6 8 10

-120

-100

-80

-60

-40

-20

0

0 50 100 150 200-60

-40

-20

0

φb=0.7 eVσ=0.1 eVVGS=-100 V

γ(tr)(1/Tn)=0.001 cm1/2V-1/2

γ(ch)(1/Tn)=0-120 V

-100 V

-80 V

-60 V

-40 V

-20 V

-10 V-5 V

Pote

ntia

l (V

)

Channel Position (μm)

VDS=-2.5 V

(V)

(nm)

Figure 4.7 (a) Our simulated surface potential profile of a representative organic transistor operating in different conditions, either in linear regime or saturation regime, upon the applied drain voltage, VDS, for a fixed gate voltage of VGS=-100 V. Here T=300 K. (b) A representative surface potential profile of the P3HT polymer transistor, as recorded in experiments by noncontact potentiometry using scanning Kelvin probe microscope by Bürgi et al. [112]., © American Institute of Physics (AIP) 2003.

94 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

To understand the peculiar surface potential drop in the vicinity of source and the

contact resistance, we recall in Section 4.2 the physics of charge injection and charge

transport as incorporated in the universal model. Apparently, both charge injection right

at the metal-organic interface and the SCLC in the short transition region cause the

substantial voltage consumption and thus significant contact resistance, as revealed by

Figure 4.7 (a) inset. We further perform simulation on the surface potential profiles for

the same transistor only with different injection barrier height (φb=0.7 eV or 0.3 eV) and

localized trap states tail width (σ=100 meV or 50 meV). Figure 4.8 shows that,

• Both charge injection barrier height and the localized trap states distribution

contributes to the surface potential profiles and thus contact resistance for organic

thin film transistors;

0 2 4 6 8 10

-80

-60

-40

-20

0

0 100 200

-40

-20

0

Pote

ntia

l (V)

Channel Position (μm)

VGS=-100 VVDS=-80 V

φb=0.7 eV, σ=0.1 eV

φb=0.3 eV, σ=0.1 eV

φb=0.3 eV, σ=0.05 eV

Pot

entia

l (V

)

Channel Position (nm)

Figure 4.8 Simulated surface potential profiles along the channel direction for the same transistor only with different injection barrier height (φb=0.7 eV or 0.3 eV) and localized trap states tail width (σ=100 meV or 50 meV). The inset is a magnified view in the 200 nm region close to the source contact.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 95

• For the same localized states distribution in the organic film, lower charge

injection barrier height at the metal-organic interface mitigates voltage drop at the

transition region and thus reduces the contact resistance. This is consistent with

the observation on output curves as shown in Figure 4.6 (a);

• Surprisingly, for the same charge injection barrier height at the metal-organic

interface, narrower localized trap states distribution, or, equivalently, smaller

Gaussian energy width of the organic film’s localized states close to the metal-

organic contacts, leads to substantially lower voltage drop in the transition region,

accordingly, giving rise to significantly reduced contact resistance. This effect is

not expected at first glance, but indeed plays a critical role in determining the

organic transistor performance as found in this study.

4.3.3 Mobility Scaling Behavior

As introduced at the beginning of this chapter, a critical issue remaining for organic

transistors is to resolve their mobility scaling behavior. By virtue of our universal

physical model in Section 4.2, we are now able to elucidate the apparently contradictory

mobility scaling features as found in experiments and reported in literature.

Figure 4.9 shows the simulation results of the mobility scaling behavior with respect

to different physical origins including carrier injection barrier height (φb), localized trap

states energy distribution (σ), Poole-Frenkel like field dependence of carrier mobility

(γ(1/Tn)), and channel length dependence of film crystallinity (γc). It is revealed that,

• Depending on the weight of the above physical effects, mobility scaling curve

with respect to the channel length (L) ranging from 100 nm to 50 µm can be

monotonically upwards, or slowly get saturated, or, surprisingly, exhibit an

overshoot region;

• Lower charge injection barrier (e.g. φb=0.3 eV vs. φb=0.7 eV), smaller localized

states distribution tail width in the transition region (e.g. σ=50 meV vs. σ=100

meV), and stronger field dependence of carrier mobility in the channel region (e.g.

96 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

γ(ch)(1/Tn))=0 vs. γ(ch)(1/Tn))=0.01 cm1/2V-1/2) all give rise to the increase of the

extrinsic field-effect mobility (μFET).

• Importantly, large modulation of the extrinsic mobility can be realized by 1-2

orders of magnitude through engineering the charge injection barrier and the

localized states distribution at the metal-organic semiconductor interface;

• The organic transistors show an extrinsic field-effect mobility increasing

monotonically with the channel length when a large injection barrier height and a

wide disordered localization states distribution occur at the contact region, which

leads to significant extrinsic contact resistance as discussed in Section 4.3.2. Such

phenomenon is consistently observed for organic transistors with the so-called

“poor contacts” in literature [76-90], and adversely compromises the benefits

from device scaling theory as introduced at the beginning of this chapter;

• It is possible to maintain a steady extrinsic field-effect mobility at a normal

channel length above the transition region width, typically, on the level of

micrometers, given that an adequate combination of charge injection barrier

height at the metal-organic interface and localized states energy distribution in the

film is established. For instance, the simulation here shows the extrinsic mobility

is almost constant for a channel length of 5-50 µm, with an injection barrier height

of 0.3 eV and a localized trap states energy distribution width of 50 meV;

• Significantly, we discover an overshoot region that may emerge in the mobility

scaling curve for organic thin film transistors within normally fabricated device

size range, provided that the electric field dependence of carrier mobility

(γ(ch)(1/Tn)) and the channel length dependence of film crystallinity (γc) in the

channel region take strong effect. This finding is understood as follows: For short

channel devices, e.g. L=100 nm to 5 µm for the simulation shown in Figure 4.9,

the extrinsic field-effect mobility is dominated by charge injection at the contact

interface and charge transport in its neighboring transition region with a low

mobility. Therefore, shrinking the channel length equivalently enhances this

dominance and leads to a lower extrinsic mobility. However, the scenario

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 97

becomes different when the channel length is far greater than the transition width,

e.g. L>5 µm for the simulation here: First, the weight of charge transport in the

channel region becomes important or even dominant; Secondly, under the same

operating voltage, shortening of the channel length increases the longitudinal

electric field strength, thus rendering a higher effective mobility in the channel

region, and as a consequence, boosting the extrinsic mobility of the device;

Additionally, for polycrystalline film transistors, the overall film crystallinity may

be enhanced for shorter channel devices [15, 27, 43], thus further increasing the

effective mobility. This effect can be quantitatively described as an extra term in

Equation (4.16) to modify the portion of free mobile states over the total states in

the channel region, Nmob(ch)(VGS, T, L)/Ntot. For simplicity, the simulation

performed here lumps together both the electric field dependence effect and the

film crystallinity effect as shown in Figure 4.9. Our finding in this study also

successfully explains the experimental results for the mobility scaling behavior of

solution-shearing processed polycrystalline 4T-TMS SPOFETs, as shown in

Figure 4.10. Note that the mobility scaling trend here is apparently opposite to

other reports [76-90], but still in agreement with our finding based on the

universal model. We emphasize that for these devices in top-contact structure,

shorter channel lengths below 10 µm are unavailable yet due to shadow mask

limitation. Future fabrication of shorter channel devices will be able to testify the

overshoot region as predicted in the model.

• The critical channel length for the peak extrinsic mobility in the overshoot region,

Lpeak, depends on the weight of the electric field dependence of carrier mobility

(γ(ch)(1/Tn)) and the channel length dependence of film crystallinity (γc) in the

channel region, the carrier injection barrier height (φb), as well as the localized trap

states energy distribution (σ). Lower injection barrier height and smaller localized

trap states energy tail width gives rises to shorter channel length where peak

extrinsic mobility appears. This suggests that for normal bottom contact devices

where the contact regions are showing poor morphology than top contact ones,

the critical channel length Lpeak can be considerably large, usually beyond the

98 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

typical device channel length as we fabricate in lab. This explains why it is rare to

find the mobility scaling trend similar to that shown in Figure 4.10 for most

bottom contact devices.

• Indeed, the critical channel length for the peak mobility serves an indication of the

contact properties, as revealed in this study.

0 10 20 30 40 50

0.00

0.05

0.10

0.15

0.20

0.25

0.1 1 100.00

0.05

0.10

0.15

0.20

Ext

rinsi

c Fi

eld-

Effe

ct M

obili

ty (c

m2 /V

s)

Channel Length (μm)

φb=0.3 eV, σ=0.05 eV, γ(tr)(1/Tn)=10-3 cm1/2V-1/2,

γ(ch)(1/Tn)+γc=10-2 cm1/2V-1/2

φb=0.3 eV, σ=0.05 eV, γ(tr)(1/Tn)=10-3 cm1/2V-1/2, γ(ch)(1/T

n)+γc=0

φb=0.3 eV, σ=0.1 eV, γ(tr)(1/Tn)=10-3 cm1/2V-1/2, γ(ch)(1/T

n)+γc=0

φb=0.7 eV, σ=0.1 eV, γ(tr)(1/Tn)=10-3 cm1/2V-1/2, γ(ch)(1/T

n)+γc=0

Lpeak

Mob

ility

(cm

2 /Vs)

Channel Length (μm)

Figure 4.9 Mobility scaling behavior with respect to different physical origins including carrier injection barrier height (φb), localized trap states energy distribution (σ), Poole-Frenkel like field dependence of carrier mobility (γ(1/Tn)), and channel length dependent film crystallinity (γc). For simplicity of the simulation here, the film crystallinity effect is lumped into the mobility’s electric field effect.

CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR 99

0 10 20 30 40 500.0

0.1

0.2

0.3

0.4

0.5

Mea

sure

d E

xtrin

sic

Mob

ility

μ FET

(cm

2 /Vs)

Channel Length, L (μm)

Lpeak

Figure 4.10 Measurement results of the extrinsic field-effect mobility with respect to different channels for the solution-shearing processed polycrystalline 4T-TMS SPOFETs. The device fabrication and measurement are detailed in Chapter 2. All of these devices are fabricated on the same wafer with the same process. The dashed line is for visual guidance and refers to the modeling and simulation results shown in Figure 4.9. It is found that the finding based on our model is in excellent agreement with the experimental results.

4.4 Summary

We have developed a universal physical model for organic transistors by

incorporating both charge injection effects at the metal-organic interface and charge

transport properties in the organic semiconductor film, and successfully applied the

device model to resolve many elusive physical phenomena observed so far, such as the

peculiar mobility scaling behavior, the contact resistance effect, and the mysterious

surface potential profiles of organic transistors which have been experimentally probed

yet poorly understood. We also discovered in this study that an overshoot region may

100 CHAPTER 4. DEVICE PHYSICS AND MODELING OF ORGANIC TRANSISTOR

emerge in the mobility scaling curve with respect to the channel length. This work

provides fresh insight into understanding of the fundamental physics of organic

transistors as well as new engineering methods for improving the organic transistor

performance.

101

Chapter 5

Contact Engineering for

Organic Electronic Device

The important thing in science is not so much to obtain new facts as to discover new ways of

thinking about them.

- William Bragg

5.1 Background

In Chapter 4, we have discussed the physics of organic thin film transistors with focus

on charge injection effects and charge transport properties. Through device modeling

and simulation in conjunction with experimental corroboration, we explicitly showed the

significant role of the contacts in determining the organic transistor performance,

including the electrical transfer and output characteristics, the surface potential profile

1 Note: Portions of this chapter are reproduced, with permission, from [45] © 2009 IEEE.

102 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

along the channel direction, the contact resistance, the extrinsic mobility and its scaling

behavior. As outlined in Chapter 4, it is crucial to lower the charge injection barrier

height and minimize the localized states energy distribution width at the metal-organic

interface to achieve optimized, high-performance devices.

Achieving this goal turns out not an easy task. It is well recognized from both

theoretical [150-152] and experimental [153-155] studies that a high Schottky injection

barrier height can emerge at the metal-organic interface irrespective of the metal work

function, a phenomenon known as Fermi-level pinning, which consequently gives rise to

poor contact properties and unfavorable organic semiconductor device performance.

In this chapter, we present the first direct evidence and demonstration of Fermi-

level depinning at metal-organic semiconductor (M/O) interfaces by inserting an

ultrathin interfacial Si3N4 insulator in between [45]. The contact behavior is successfully

tuned from rectifying to quasi-Ohmic and to tunneling by varying the Si3N4 thickness

within 0-6 nm. Detailed physical mechanisms of Fermi-level pinning/depinning

responsible for the M/O contact behavior are clarified based on a proposed lumped-

dipole model. Experimental results are in good agreement with the theory and the

model.

5.2 Experimental Results on Fermi-Level Depinning

We chose p- and n-type organic semiconductors, pentacene and 3,4,9,10-perylene-

tetracarboxylic-dianhydride (PTCDA), respectively, for this study. Both materials have

been found to exhibit relatively strong Fermi-level pinning at their M/O interfaces [156],

with the pinning factor (S=dφb/dψmetal) of S~0.4 for metal/pentacene and S~0 for

metal/PTCDA. Si3N4 was chosen as the interfacial layer because (1) less-oxygen ambient

suppresses unintentional oxidation of the underlying material, and (2) Si3N4 has a

moderately large bandgap with its conduction/valence band relatively far away from the

LUMO/HOMO level of the aforementioned organic semiconductors, thus excluding

the possibility of charge injection assisted by a close transition energy level as

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 103

traditionally suggested for organic LEDs or solar cells. Si3N4 with different thicknesses

of 0-6 nm (measured from dummy silicon wafers by ellipsometry) was deposited at

room temperature by a precisely-controlled high-vacuum LSI sputtering system.

Figure 5.1 illustrates the device structure and the process flow for our

Au/Si3N4/pentacene diodes. Atomic force microscope (AFM) images of the pentacene

film as shown in Figure 5.2 indicate that there is no damage caused by the Si3N4

sputtering process on the underlying organic semiconductor layer. A large Au electrode

is used as the common anode, and the cathode electrodes with a much smaller area are

either Au pads deposited by thermal evaporation or Au bonding wires. We found that

the diodes with/without Si3N4 based on the evaporated Au pads are regularly shorted

due to Au atoms penetration, leading to an artifactual Ohmic contact with very low

resistance as shown in Figure 5.3.

Figure 5.1 Device structure and process flow of the Au/Si3N4/pentacene diodes. Pentacene was chosen for this study since the pinning factor (S=dφb/dψmetal) S~0.4 is relatively small for the metal/pentacene interface.

104 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

Figure 5.2 AFM images of 200 nm pentacene films deposited on Au at room temperature (a) before and (b) after Si3N4 sputtering deposition. Height scale: 300 nm.

-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6

-200

-100

0

100

200

Curre

nt d

ensit

y (A/

cm2 )

Voltage (V)

t(SiN)=4.1 nm t(SiN)=2.9 nm t(SiN)=1.1 nm

Figure 5.3 I-V measurement results showing the diodes with/without Si3N4 based on evaporated Au pads are regularly shorted due to metal atoms penetration along the grain boundaries or the roughness-induced vacancies.

To resolve the shorting issues and eliminate the electrical artifacts resulting from the

gold atom penetration, we introduced the gold wire based diode configurations for this

study as depicted in Figure 5.1. Figure 5.4 shows the measured I-V characteristics of the

Au/Si3N4/pentacene diodes with different Si3N4 thicknesses, providing direct evidence

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 105

that the metal-organic semiconductor diodes are successfully tuned from rectifying to

quasi-Ohmic and to tunneling ones with increase of the Si3N4 thickness in between, thus

in corroboration with the Fermi-level pinning/depinning theory and model as

introduced in following section. Quasi-Ohmic contact behavior for Au/pentacene is

observed when the Si3N4 thickness is nominally ~4 nm. Figure 5.5 shows the

normalized dynamic resistance (RAC=∂V/∂I) and static resistance (RDC=V/I) of the

Au/Si3N4/pentacene diodes with respect to different Si3N4 thickness. Both resistances

reach the minimum at an optimal Si3N4 thickness of ~4.1 nm, a further evidence of

effective Fermi-level depinning and contact resistance reduction by the insertion of the

Si3N4 interfacial layer. At first glance it is surprising that a nominal Si3N4 thickness of ~4

nm leads to quasi-Ohmic contacts from the tunneling theory point of view. After taking

into account surface roughness of the underlying layer which lowers the effective Si3N4

thickness, we argue that such an experimental observation is still expectable, and in no

doubt is particularly important for practical applications.

106 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

-1.0 -0.5 0.0 0.5 1.0-150-100-50

050

100150

t(SiN)=4.1 nmAu/SiN/Pentacene

Voltage (V)

Curre

nt (n

A)

-1.0 -0.5 0.0 0.5 1.0-10

0

10

20

30

40

Curre

nt (n

A)

Voltage (V)

no SiN:Au/Pentacene

-1.0 -0.5 0.0 0.5 1.0-0.6-0.4-0.20.00.20.40.6

t(SiN)=6.1 nmAu/SiN/Pentacene

Voltage (V)

Curre

nt (n

A)

Figure 5.4 I-V characteristics of the Au/Si3N4/pentacene diodes with different Si3N4 thicknesses (with Au wire as the top cathode electrode), providing direct evidence that the Au/pentacene diode has been successfully tuned to rectifying, quasi-Ohmic, and symmetric tunneling behavior by modulating the Si3N4 thickness. (The effective contact area between Au wire and Si3N4/pentacene is only a fraction of the wire cross section due to the surface roughness of Si3N4/pentacene layer, and may vary from device to device. One should note that the I-V shape giving the diode property is the essence here.)

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 107

-1.0 -0.5 0.0 0.5 1.0105

106

107

108

109

1010

1011

1012

Stat

ic re

sista

nce (

arb.

uni

t)

Voltage (V)

RDC=V/I

-1.0 -0.5 0.0 0.5 1.0105

106

107

108

109

1010

1011

t(SiN)=0 nm t(SiN)=1.1 nm t(SiN)=1.8 nm t(SiN)=2.9 nm t(SiN)=4.1 nm t(SiN)=6.1 nm

RAC=dV/dI

Dyna

mic

resis

tanc

e (ar

b. u

nit)

Voltage (V)

(b)b

a

Figure 5.5 Normalized dynamic resistance (RAC=∂V/∂I) and static resistance (RDC=V/I) of the Au/Si3N4/pentacene diodes with respect to the Si3N4 thickness, as calculated from their respective I-V measurement curves. Normalization is based on the fact that the maximum forward-biased diode current (here at V=+1 V) is less affected by the Si3N4 thickness. Note: The device with a thick Si3N4 layer (t=6.1 nm) was not normalized here as tunneling dominates therein.

108 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

We also fabricated Ag/Si3N4/PTCDA diode array as shown in Figure 5.6 for this

study. Instead of measuring an individual diode’s I-V curve, here we evaluated the

contact resistance of Ag/Si3N4/PTCDA by measuring two back-to-back diodes in series

connection with a low-conductivity PTCDA bulk film [see Figure 5.7 (a)]. Figure 5.7 (b)

shows explicitly a minimum contact resistance observed for the Ag/PTCDA interface

with an optimal sandwiched Si3N4 thickness of ~2.5 nm, reaffirming the Fermi-level

depinning at M/O interfaces by inserting the ultrathin Si3N4 interfacial layer. The

apparent difference of the optimal Si3N4 thickness for Au/pentacene and Ag/PTCDA

interfaces can be attributed to their different interface roughness (and thus different

effective interfacial layer thickness). This also underscores that one should consider the

particular device structure and materials therein when applying the Fermi-level

depinning technology to high performance organic devices.

(A) (A)(B)

(C)

(D)

(A) (A)

(A) Ag electrode array with thickness=30 nm and spacing=250 μm acting as bottom electrodes; (B) Si3N4 layer with different thicknesses of 0.6 nm, 1.0 nm, 1.9 nm, 2.4 nm, 3.2 nm, 3.9 nm, 6.1 nm; (C) 130 nm PTCDA deposited at room temperature by thermal evaporation;(D) 300 nm SiO2 wafer as the device substrate (holder).

O

OO

O

OO

PTCDA

(A) (B) (C)

1mm

Figure 5.6 Device structure, microscopy image, and process flow of the Ag/Si3N4/PTCDA diodes. PTCDA was chosen here since the pinning factor (S=dφb/dψmetal) is S~0 for metal/PTCDA interface, indicating very strong Fermi-level pinning effect at the metal/PTCDA interface.

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 109

0 ( )

0

0

( )

0( )

,

.

;

;

total c sh film

c

cc

film sh film

cctotal film sh film

R R R l W

R Const

RR Area W LW L

lR RL

R lR R R RW L L

W L α

α

α

⇓⋅ ≈ + ⋅ ⋅ ⋅

≈ +

= = ⋅⋅

= ⋅

≈ + = + ⋅⋅

0 1 2 3 4

108

109

-15

-10

-5

0

To

tal r

esist

ance

(Ω cm

2 )

Si3N4 Thickness (nm)

Norm

alize

d Re

lative

Co

ntac

t Res

istan

ce Rc0

(arb

. uni

t)

Figure 5.7 (a) A quick and straightforward extraction method for the relative contact resistance of the diodes with different Si3N4 thicknesses, through measurement of the I-V characteristics on two adjacent electrodes; (b) A minimum contact resistance is observed at the Ag/PTCDA interface with a sandwiched Si3N4 thickness of ~2.5 nm, reaffirming the Fermi-level depinning at M/O interfaces by inserting the ultrathin interfacial Si3N4 layer.

5.3 Theory and Proposed Model

Intuitively, introducing an ultrathin Si3N4 layer in between the metal-organic

semiconductor interface may give rise to a few effects. One is modulation of the organic

film’s localized states distribution close to the interface, i.e. σ as discussed in Chapter 4,

since the Si3N4 layer either protects the film from physical interference of metal

110 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

deposition for top-contact structure (Figure 5.1) or facilitates better film growth on the

metal for bottom-contact structure (Figure 5.6). The other one is modulation of the

charge injection barrier height, i.e. φb. We now proceed to discuss the mechanisms

responsible for this effect, Fermi-level depinning, as briefly mentioned at the beginning

of this chapter.

Figure 5.8 Different mechanisms of Fermi-level pinning at the metal/organic semiconductor interface, including (a) MIGS and intrinsic surface states, (b) charge transfer, (c) covalent bonding, (d) permanent molecular dipole, (e) image force, and (f) exchange/Pauli repulsion. An interface dipole is always created upon the M/O junction formation.

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 111

Classical metal-induced gap states (MIGS) theory [157] has been applied to explain

Fermi-level pinning/depinning at traditional metal-inorganic semiconductor (e.g. Si [158]

or Ge [159]) interface. The concept is simply illustrated in Figure 5.8 (a): free-electron

wave function penetrates into the semiconductor bandgap, giving rise to a large amount

of MIGS which pin the Fermi energy close to the charge neutrality level (CNL) and

form a large Schottky barrier for charge injection. However, the scenario for M/O

interfaces is found more complicated, and a variety of mechanisms as depicted in Figure

5.8 (a)-(f) have been suggested to understand the M/O interfacial electronic structures

[154-156]. The overall effect is an interface dipole (Δ) created upon the formation of the

M/O junction. Note that there can be disorder-induced energy level fluctuations and

Anderson localized states at the M/O interface as discussed in Chapter 4, which are

shown only leading to an effective barrier lowering by σ2/2kT. For simplicity and clarity,

effective HOMO/LUMO levels are denoted in this chapter.

Figure 5.9 (a)-(b) shows a generalized picture of the M/O interface energy band

diagram, indicating that the Fermi-level pinning manifests itself as formation of a

lumped interface dipole with its magnitude depending on the metal. The measured

injection barrier height φb(real) can be considerably larger than the ideal φb(ideal) following

Schottky–Mott limit due to the interface dipole induced energy shift ∆0, and can be

modeled by

0

00

cos

b real b ideal metal metal

i i

i r

U HOMO HOMO

N P dπ

φ φ ψ ψ

θ θ θ θε ε

= + Δ = − ⋅ −

⋅ ⋅+∑∫ (5.1)

where U(x) is the unit step function, θ is the alignment angle between the dipole element

and the interface normal, Ni and Pi are, respectively, the density and moment of the

dipole element type i, corresponding to different dipole origins as shown in Figure 5.8.

Since large charge injection barrier at the source/drain of organic TFTs leads to

significant contact resistance, deteriorating TFT electrical performance and their

112 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

mobility scaling behavior as discussed in Chapter 4, it is crucial to mitigate Fermi-level

pinning effect at such M/O interfaces, and the dipole elements from all aforementioned

mechanisms, Ni and Pi, should be eliminated.

00

( ) ( ) cos /( )i i ri

N P dπ

θ θ θ θ ε εΔ = ∑∫

Figure 5.9 M/O interface energy band diagrams based on the proposed lumped interface dipole model: (a) before M/O interface formation; (b) upon M/O interface formation, Fermi-level pinning arises from various interface dipole elements; (c) after inserting an ultrathin Si3N4 insulator, the Fermi-level depinning takes effect by blocking the physisorption/chemisorption.

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 113

As illustrated in Figure 5.9 (c), an ideal insulator shields

physisorption/chemisorption interaction at the M/O interface and thus releases the

Fermi level pinning. A quasi-zero Schottky barrier or Ohmic contact is therefore

expected under depinning circumstances.

To more quantitatively understand the Fermi-level depinning effect, we propose a

simple depinning model based on the fact that there are two competing mechanisms

responsible for the charge injection current or contact resistance Rc after inserting the

thin insulator. One is direct or Fowler-Nordheim (FN) tunneling through the insulator

with its probability exponentially decreasing with insulator thickness t. The other is

charge injection over the effective Schottky barrier φb(real) which decreases with the

insulator thickness (t) due to the depinning effect. Therefore, assuming roughly that

( ) ( ) 0 exp( / )nb real b ideal ct tφ φ= + Δ ⋅ − (5.2)

where n>1, tc is a critical thickness parameter, the contact resistance thus can be

described as

0 1 ( ) 0 2exp[ ( exp( / ) )] exp( )nc c b ideal cR R t t tβ φ β≅ ⋅ ⋅ + Δ ⋅ − ⋅ ⋅ (5.3)

where β1 and β2 are thermionic emission relevant coefficient and tunneling relevant

coefficient, respectively.

Figure 5.10 shows an exemplified simulation plot for Rc-t based on this depinning

model, which is in good agreement with our experimental results as presented in last

section [see Figure 5.5 and Figure 5.7]. It is also found that the insulator thickness must

be optimized carefully: initial increase of the insulator thickness reduces contact

resistance due to Schottky barrier modulation; once beyond the optimal thickness,

tunneling resistance through the insulator becomes dominated rapidly.

114 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

Figure 5.10 Simulation results based on the proposed simple depinning model showing the dominant mechanisms responsible for the contact resistance behavior after insertion of the ultrathin insulating layer at the M/O interface.

5.4 Summary

We have successfully demonstrated Fermi-level depinning at two different M/O

interfaces by inserting a precisely-controlled ultrathin interfacial Si3N4 layer. The contact

behavior is tuned from rectifying to quasi-Ohmic and to tunneling by modulation of the

CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE 115

Si3N4 thickness within 0-6 nm. We also proposed a lumped dipole model to clarify the

detailed physical mechanisms of Fermi-level pinning/depinning at the M/O interface.

Experimental results are in good agreement with the theory and the model. This work

represents a significant step toward the fundamental understanding of M/O interface

properties and technological advancement of achieving low-resistance Ohmic contacts

for organic electronic device applications, and can be particularly useful for optimization

of the organic transistor performance through source/drain contact engineering.

116 CHAPTER 5. CONTACT ENGINEERING FOR ORGANIC ELECTRONIC DEVICE

117

Chapter 6

Conclusions and Outlook

Give me a fulcrum, and I shall move the world.

- Archimedes of Syracuse (Ἀρχιμήδης)

In this dissertation, we have systematically studied carbon or organic semiconductor

thin-film field-effect transistors, particularly, solution-processed organic field-effect

transistors (SPOFETs), toward application in the emerging flexible electronics and

macroelectronics fields. The topics described span from fundamental device physics,

device modeling, fabrication technology to interface/contact engineering, highlighting

the most important aspects for the development of organic transistor based electronic

devices and systems.

6.1 Conclusions

Revisit the dissertation, there are a number of research contributions originated from

this work. A brief summary of the contributions here serves as the conclusions of this

dissertation.

118 CHAPTER 6. CONCLUSIONS AND OUTLOOK

• Designed a novel device architecture for flexible SPOFETs; Demonstrated and

characterized high performance flexible SPOFETs on plastic with a carrier

mobility over 0.2 cm2/Vs, a turn-on voltage of near 0 V, and a record low

subthreshold slope of ~80 mV/dec in ambient conditions. These exceptional

characteristics are achieved by excellent electrostatic control, 3-D statistical

modeling for solution-shearing process optimization, and phenyl-terminated self-

assembled monolayer (SAM) based interface engineering.

• Utilized 3-D statistical modeling and data analysis to optimize the solution-

shearing process; Provided a general guideline of process optimization for high

performance organic transistors.

• Systematically investigated dipole moment related physical effects and chemistry

effects of the SAM at the organic semiconductor-dielectric interface; Elucidated

how the performance and reliability of organic transistors are controlled by their

interface conditions through careful selection of a group of phenyl-terminated

SAMs; Proposed electric shielding tensors effect induced by SAM dipole as a new

physical mechanism involved in the performance control of organic transistors.

• Introduced a simple spin-coating process for deposition of high-quality phenyl-

terminated SAMs for organic electronics applications.

• Proposed and developed a universal physical model and 1-D device simulator for

organic transistors by incorporating both charge injection effects at the metal-

organic interface and charge transport properties in the organic semiconductor

film; Successfully applied the device model to resolve many elusive physical

phenomena observed so far, such as the peculiar mobility scaling behavior, the

contact resistance effect, and the mysterious surface potential profiles along the

channel which have been experimentally probed yet poorly understood; Validated

the model by excellent agreement between the simulation results and the

experimental results.

• Systematically elucidated the device physics involved in the charge injection and

charge transport in organic semiconductor thin-film transistors; Discovered an

CHAPTER 6. CONCLUSIONS AND OUTLOOK 119

overshoot region in the mobility scaling behavior and identified the existence of a

critical channel length for the peak field-effect mobility.

• Explored and demonstrated Fermi-level depinning at the metal-organic interface

for low-resistance Ohmic contacts by inserting an ultrathin interfacial Si3N4

insulator in between, which lowers charge injection barrier and reduces the

interfacial disorder width or localization states; Successfully tuned the contact

behavior from rectifying to quasi-Ohmic and to tunneling by varying the Si3N4

thickness within 0-6 nm;

• Clarified the detailed physical mechanisms of Fermi-level pinning/depinning

responsible for the metal-organic semiconductor contact behavior based on a

proposed lumped-dipole model.

6.2 Outlook

There is never an end to the evolution of science and technology itself. Despite the fact

that encouraging progress has been achieved as of today in the development of organic

semiconductor transistors, many challenges and thus opportunities remain in future

studies. For the work presented in this dissertation, we address the relevant future

directions as follows.

• For the solution-shearing process described in Chapter 2, we have performed

systematical experiment design and 3-D statistical data modeling/analysis to find

out the optimal processing conditions. Indeed, it is possible to develop an

analytical model for this process and numerically simulate the growth output

based on the evaporation kinetics and crystallization kinetics. This would require

more knowledge of parameter values for the involved materials and environments.

• Based on the high-performance flexible SPOFETs introduced in Chapter 2, it is

feasible to demonstrate flexible integrated circuits and systems and bring the

flexible organic transistors closer to practical applications. A particularly

120 CHAPTER 6. CONCLUSIONS AND OUTLOOK

interesting design would be integration of the SPOFET and OLED on a single

flexible substrate for all-organic displays.

• As briefly mentioned at the end of Chapter 3, there is an important physical effect

of the SAM at the dielectric-semiconductor interface has not been considered yet

to date, i.e. strong external interfacial electric field as induced by the gate voltage

would distort the inherent dipole moment of the SAM molecule as illustrated in

Figure 3.19, thus influence the performance and reliability of organic transistors.

This effect, namely electric shielding tensors effect, must be given careful

consideration in future studies. From the experiment point of view, in-situ and

real-time measurement of the SAM dipole on an oxide surface under the applied

external electric field would testify and give direct evidence of the electric

shielding tensors effect. In the meantime, it is also possible to model this effect

using Ab-initio simulation.

• Finally, it is possible to upgrade current 1-D universal device model and device

simulator as introduced in Chapter 4 to a 2-D or 3-D simulator in the future.

121

Appendix A

Spin-coating Process for Phenyl-

Terminated Self-Assembled

Monolayer

A.1 Summary

Use simple spin-coating method to deposit high-quality, densely packed phenyl-

terminated self-assembled monolayers (SAM) on various oxides or OH-terminated

polymeric dielectric surfaces. The resultant SAMs can be used for organic electronic

devices including OFETs and diodes (particularly useful for solution-processed organic

electronics since the phenyl-SAMs possess ideal tradeoff between surface energy and

wettability for organic semiconductor solution). The electrical performance of these

devices can be superior to those based on the same phenyl-SAMs deposited by

traditional methods such as soaking.

122 APPENDIX A. SPIN-COATING PROCESS FOR PHENYL-TERMINATED SAM

A.2 Materials

(1) Phenyl-terminated silanes, such as Phenyltrichlorosilane (PTS),

Phenethyltrichlorosilane (PETS), 4-Phenylbutyltrichlorosilane (PBTS), 6-

Phenylhexyltrichlorosilane (PHTS), N-Phenylaminopropyltricholorosilane (PAPTS).

[Note: In this work, these materials were ordered from Gelest, Inc.]

(2) Anhydrous toluene

A.3 Tools

(1) Inert environment such as N2 glovebox or glovebag

(2) Spin-coater in ambient conditions

(3) Accessories such as pipettes, tips, syringes and vials

A.4 Substrate Preparation

(1) Prepare and clean substrates: for Si/SiO2 substrate, use piranha treatment (7:3

mixture of H2SO4 and H2O2, highly corrosive and oxidizing);

(2) Pre-treat substrate right before spin-coating: for piranha treated Si/SiO2

substrates, put them in UV-Ozone for 30 mins to 1 hr.

A.5 Phenyl-SAM Solution Preparation

(1) Transfer phenyl-silanes and anhydrous toluene to an inert environment, such as

N2 glovebox or glovebag; other accessories such as pipettes, tips, syringes and vials

should be transferred in all together.

APPENDIX A. SPIN-COATING PROCESS FOR PHENYL-TERMINATED SAM 123

[Note: the phenyl-silane solution in their original stock bottles should be completely clear otherwise it

has been probably polymerized. Please check this carefully. Polymerization of the original phenyl-silane

solution could significantly affect the final SAM quality]

(2) Prepare phenyl-SAM solution in anhydrous toluene in the inert environment: the

phenyl-SAM solution concentration is typically 1-1.5:1000=phenyl-silane:anhydrous

toluene by volume.

(3) Start by filtering the anhydrous toluene through a 0.2 µM filter into a 20 ml vial;

transfer 1-2 ml phenyl-silane solution from their original stock bottle into another vial

and close the original stock bottle lid immediately to minimize contamination for future

use. Prepare the phenyl-SAM solution with the appropriate concentration indicated

above from those vials.

(4) Transfer the prepared SAM solution in anhydrous toluene from the inert

glovebox/glovebag to outside.

A.6 Spin-coating for Phenyl-SAM Devices

(1) Prepare the spin-coating setup: Clean the spin-coater holder with appropriate

solvents; set the spin-rate to 3 Krpm, acceleration rate to 0.

(2) Spin-coat phenyl-SAMs on the device substrates: place the device substrate

prepared in “step-[A.4]” on the spin-coater holder; use pipette to place the phenyl-SAM

solution prepared in “step-[A.5]” onto the device substrate; make sure the solution

cover the whole substrate surface completely; wait for 10-15 secs; start spin-coating at 3

Krm for another 45-50 secs.

[Note: Please try to make the solution cover the whole substrate surface by once to minimize extra

polymerization in air. Typically a 150 uL volume of SAM solution should be used for a substrate size

of 1-4 cm2. Immediately recap the solution after each drop]

124 APPENDIX A. SPIN-COATING PROCESS FOR PHENYL-TERMINATED SAM

A.7 Post-processing

(1) Place the device substrates after spin-coating of the phenyl-SAMs in a large

desiccator or brown glass bottle. Make sure the desiccator or brown glass bottle are

clean inside to avoid contamination to the wafer backside. The contamination can bring

defects to the SAM surface during the subsequent sonication process.

(2) Prepare ~1 mL of Hydrochloric Acid (HCl) into a small 2 mL vial; place this 2

mL vial with HCl into a 20 mL vial that has been half-full with the small silica particles

used in column purification; leave both vials uncapped; place these vials into the large

desiccator or brown bottle which have samples inside; tightly close lid to the descicator

or brown bottle. Leave them undisturbed for 24-48 hrs, or at least overnight.

(3) Open the desiccator or brown bottle in the hood for safety. Remove samples and

copiously rinse with HCLP toluene; place samples in glass-ware and sonicate for 5 mins

TWICE with each solvent in the following order: (HCLP-level) Toluene, Acetone,

Isopropyl Alcohol.; dry clean with N2 gun.

(4) Store samples. Rinse the samples with the solvents and dry with N2 every time

before the use.

A.8 Reference

Y. Ito, et al., "Crystalline Ultrasmooth Self-Assembled Monolayers of Alkylsilanes

for Organic Field-Effect Transistors," Journal of the American Chemical Society, vol. 131, pp.

9396-9404, Jul 8 2009.

H. Y. Nie, et al., "Delivering octadecylphosphonic acid self-assembled monolayers

on a Si wafer and other oxide surfaces," Journal of Physical Chemistry B, vol. 110, pp.

21101-21108, Oct 26 2006.

125

Bibliography

[1] D. J. Frank, et al., "Device scaling limits of Si MOSFETs and their application

dependencies," Proceedings of the IEEE, vol. 89, pp. 259-288, Mar 2001.

[2] P. Avouris, et al., "Carbon-based electronics," Nature Nanotechnology, vol. 2, pp.

605-615, 2007.

[3] T. Durkop, et al., "Extraordinary mobility in semiconducting carbon nanotubes,"

Nano Letters, vol. 4, pp. 35-39, 2003.

[4] Y.-M. Lin, et al., "Operation of graphene transistors at gigahertz frequencies,"

Nano Letters, vol. 9, pp. 422-426, 2008.

[5] Z. Chen, et al., "An integrated logic circuit assembled on a single carbon

nanotube," Science, vol. 311, p. 1735, March 2006.

[6] Z. Bao and J. Locklin (Eds.), Organic Field-Effect Transistors. Boca Raton, FL: CRC

Press, Taylor & Francis Group, 2007.

[7] R. A. Street, "Thin-film transistors," Advanced Materials, vol. 21, pp. 2007-2022,

May 2009.

[8] S. R. Forrest, "The path to ubiquitous and low-cost organic electronic appliances

on plastic," Nature, vol. 428, pp. 911-918, Apr 2004.

[9] S. Liu, et al., "Controlled deposition of crystalline organic semiconductors for

field-effect-transistor applications," Advanced Materials, vol. 21, pp. 1217-1232,

2009.

[10] G. E. Jabbour, et al., "Screen printing for the fabrication of organic light-emitting

devices," IEEE Journal of Selected Topics in Quantum Electronics, vol. 7, pp. 769-773,

2001.

126 BIBLIOGRAPHY

[11] T. Sekitani, et al., "Organic transistors manufactured using inkjet technology with

subfemtoliter accuracy," Proceedings of the National Academy of Sciences, vol. 105, pp.

4976-4980, Apr 2008.

[12] H. Sirringhaus, et al., "High-resolution inkjet printing of all-polymer transistor

circuits," Science, vol. 290, pp. 2123-2126, Dec 2000.

[13] V. Subramanian, et al., "Progress Toward Development of All-Printed RFID

Tags: Materials, Processes, and Devices," Proceedings of the IEEE, vol. 93, pp.

1330-1338, 2005.

[14] J. A. Rogers, et al., "Paper-like electronic displays: Large-area rubber-stamped

plastic sheets of electronics and microencapsulated electrophoretic inks,"

Proceedings of the National Academy of Sciences of the United States of America, vol. 98,

pp. 4835-4840, Apr 2001.

[15] Z. Liu, et al., "Experimental study and statistical analysis of solution-shearing

processed organic transistors based on an asymmetric small-molecule

semiconductor," IEEE Transactions on Electron Devices, vol. 56, pp. 176-185, Feb

2009.

[16] D. H. Kim, et al., "High-mobility organic transistors based on single-crystalline

microribbons of triisopropylsilylethynyl pentacene via solution-phase self-

assembly," Advanced Materials, vol. 19, pp. 678-682, 2007.

[17] S. C. B. Mannsfeld, et al., "Highly efficient patterning of organic single-crystal

transistors from the solution phase," Advanced Materials, vol. 20, pp. 4044-4048,

2008.

[18] H. Klauk, "Organic circuits on flexible substrates," in Technical Digest of IEEE

International Electron Devices Meeting (IEDM), 2005, pp. 446-449.

[19] H. Sirringhaus, "Device Physics of Solution-Processed Organic Field-Effect

Transistors," Advanced Materials, vol. 17, pp. 2411-2425, 2005.

BIBLIOGRAPHY 127

[20] G. H. Gelinck, et al., "High-performance all-polymer integrated circuits," Applied

Physics Letters, vol. 77, pp. 1487-1489, 2000.

[21] J. A. Rogers, et al., "Organic smart pixels and complementary inverter circuits

formed on plastic substrates by casting and rubber stamping," IEEE Electron

Device Letters, vol. 21, pp. 100-103, 2000.

[22] L. Torsi and A. Dodabalapur, "Organic Thin-Film Transistors as Plastic

Analytical Sensors," Analytical Chemistry, vol. 77, pp. 380 A-387 A, 2005.

[23] G. H. Gelinck, et al., "Flexible active-matrix displays and shift registers based on

solution-processed organic transistors," Nature Materials, vol. 3, pp. 106-110,

2004.

[24] R. A. Street, et al., "Jet printing flexible displays," Materials Today, vol. 9, pp. 32-

37, 2006.

[25] R. Blache, et al., "Organic CMOS circuits for RFID applications," in Digest of

Technical Papers, IEEE International Solid-State Circuits Conference (ISSCC), 2009, pp.

208-209.

[26] K. Myny, et al., "A 128b organic RFID transponder chip, including Manchester

encoding and ALOHA anti-collision protocol, operating with a data rate of

1529b/s," in Digest of Technical Papers, IEEE International Solid-State Circuits

Conference (ISSCC), 2009, pp. 206-207.

[27] D. J. Gundlach, et al., "Contact-induced crystallinity for high-performance

soluble acene-based transistors and circuits," Nature Materials, vol. 7, pp. 216-221,

2008.

[28] S. H. Lee, et al., "High-performance thin-film transistor with 6,13-

bis(triisopropylsilylethynyl) pentacene by inkjet printing," Organic Electronics, vol.

9, pp. 721-726, 2008.

[29] J. H. Cho, et al., "Printable ion-gel gate dielectrics for low-voltage polymer thin-

film transistors on plastic," Nature Materials, vol. 7, pp. 900-906, 2008.

128 BIBLIOGRAPHY

[30] H. Yan, et al., "A high-mobility electron-transporting polymer for printed

transistors," Nature, vol. 457, pp. 679-686, 2009.

[31] S. Scheinert, et al., "Subthreshold characteristics of field effect transistors based

on poly(3-dodecylthiophene) and an organic insulator," Journal of Applied Physics,

vol. 92, pp. 330-337, 2002.

[32] T. Someya, et al., "A large-area, flexible pressure sensor matrix with organic field-

effect transistors for artificial skin applications," Proceedings of the National Academy

of Sciences of the United States of America, vol. 101, pp. 9966-9970, Jul 2004.

[33] R. Parashkov, et al., "All-organic thin-film transistors made of poly(3-

butylthiophene) semiconducting and various polymeric insulating layers," Journal

of Applied Physics, vol. 95, pp. 1594-1596, 2004.

[34] Z. Bao, et al., "Silsesquioxane Resins as High-Performance Solution Processible

Dielectric Materials for Organic Transistor Applications," Advanced Functional

Materials, vol. 12, pp. 526-531, 2002.

[35] M. E. Roberts, et al., "Water-stable organic transistors and their application in

chemical and biological sensors," Proceedings of the National Academy of Sciences of the

United States of America, vol. 105, pp. 12134-12139, Aug 2008.

[36] Z. Liu, et al., "Solution-processed flexible organic transistors showing very-low

subthreshold slope with a bilayer polymeric dielectric on plastic," Applied Physics

Letters, vol. 94, p. 203301, May 2009.

[37] Z. Liu, et al., "High-Performance Air-Stable Solution Processed Organic

Transistors," in Proceedings of 66th IEEE Annual Device Research Conference (DRC),

2008, pp. 239-240.

[38] H. A. Becerril, et al., "High-performance organic thin-film transistors through

solution-sheared deposition of small-molecule organic semiconductors,"

Advanced Materials, vol. 20, pp. 2588-2594, Jul 2008.

BIBLIOGRAPHY 129

[39] H. A. Becerril, et al., "Solution Sheared Deposition of Organic Semiconductor

Thin-Films with Oriented Crystalline Morphology," U.S. patent on file, Docket

No.:S08-142 at Stanford Office of Technology Licensing Patent, 2008.

[40] C. J. Tonzola, et al., "New n-Type Organic Semiconductors: Synthesis, Single

Crystal Structures, Cyclic Voltammetry, Photophysics, Electron Transport, and

Electroluminescence of a Series of Diphenylanthrazolines," Journal of the American

Chemical Society, vol. 125, pp. 13548-13558, 2003.

[41] Y.-Y. Lin, et al., "Pentacene-based organic thin-film transistors," IEEE

Transactions on Electron Devices, vol. 44, pp. 1325-1331, 1997.

[42] A. Virkar, et al., "The role of OTS density on pentacene and C60 nucleation, thin

film growth, and transistor performance," Advanced Functional Materials, vol. 19,

pp. 1962-1970, 2009.

[43] Z. Liu, et al., "High-mobility Air-stable Solution-Shear-Processed Organic

Transistors based on an Asymmetric Small Molecule for Flexible Electronic

Applications," Materials Research Society (MRS) Spring Meeting, San Francisco, CA,

Mar 24-28 2008.

[44] C.-C. Kuo, et al., "TES anthradithiophene solution-processed OTFTs with 1

cm^2/Vs mobility," in Technical Digest of IEEE International Electron Devices Meeting

(IEDM), 2004, pp. 373-376.

[45] Z. Liu, et al., "Fermi Level Depinning at Metal-Organic Semiconductor Interface

for Low-Resistance Ohmic Contacts," Technical Digest of IEEE International

Electron Devices Meeting (IEDM), 2009.

[46] V. Erukhimovitch and J. Baram, "Crystallization kinetics," Physical Review B, vol.

50, p. 5854, 1994.

[47] H. Xue, et al., "Experimental Studies of Droplet Evaporation Kinetics:

Validation of Models for Binary and Ternary Aqueous Solutions," Journal of the

Atmospheric Sciences, vol. 62, pp. 4310-4326, 2005.

130 BIBLIOGRAPHY

[48] "IEEE standard test methods for the characterization of organic transistors and

materials," IEEE Std 1620-2004, 2004.

[49] T. B. Singh and N. S. Sariciftci, "Progress in Plastic Electronics Devices," Annual

Review of Materials Research, vol. 36, pp. 199-230, 2006.

[50] Y. Shirota and H. Kageyama, "Charge Carrier Transporting Molecular Materials

and Their Applications in Devices," Chemical Reviews, vol. 107, pp. 953-1010,

2007.

[51] J. E. Anthony, "Functionalized Acenes and Heteroacenes for Organic

Electronics," Chemical Reviews, vol. 106, pp. 5028-5048, 2006.

[52] A. Salleo, "Charge transport in polymeric transistors," Materials Today, vol. 10, pp.

38-45, Mar 2007.

[53] V. Coropceanu, et al., "Charge transport in organic semiconductors," Chemical

Reviews, vol. 107, pp. 926-952, Apr 2007.

[54] T. M. Brown, et al., "Built-in field electroabsorption spectroscopy of polymer

light-emitting diodes incorporating a doped poly(3,4-ethylene dioxythiophene)

hole injection layer," Applied Physics Letters, vol. 75, pp. 1679-1681, 1999.

[55] H. Klauk, et al., "Ultralow-power organic complementary circuits," Nature, vol.

445, pp. 745-748, 2007.

[56] A. Rolland, et al., "Electrical Properties of Amorphous Silicon Transistors and

MIS-Devices: Comparative Study of Top Nitride and Bottom Nitride

Configurations," Journal of the Electrochemical Society, vol. 140, pp. 3679-3683, 1993.

[57] A. Silvia, et al., "Water-Soluble, Electroactive, and Photoluminescent

Quaterthiophene-Dinucleotide Conjugates," Chemistry - A European Journal, vol.

14, pp. 513-521, 2008.

[58] W. F. Brinkman, et al., "A history of the invention of the transistor and where it

will lead us," IEEE Journal of Solid-State Circuits, vol. 32, pp. 1858-1865, 1997.

BIBLIOGRAPHY 131

[59] Y. Y. Lin, et al., "Stacked pentacene layer organic thin-film transistors with

improved characteristics," IEEE Electron Device Letters, vol. 18, pp. 606-608, 1997.

[60] K. Fukuda, et al., "Effects of the alkyl chain length in phosphonic acid self-

assembled monolayer gate dielectrics on the performance and stability of low-

voltage organic thin-film transistors," Applied Physics Letters, vol. 95, p. 203301,

2009.

[61] D. Kumaki, et al., "Significant improvement of electron mobility in organic thin-

film transistors based on thiazolothiazole derivative by employing self-assembled

monolayer," Applied Physics Letters, vol. 90, p. 053506, 2007.

[62] Y. Wu, et al., "Controlled orientation of liquid-crystalline polythiophene

semiconductors for high-performance organic thin-film transistors," Applied

Physics Letters, vol. 86, p. 142102, 2005.

[63] K. P. Pernstich, et al., "Threshold voltage shift in organic field effect transistors

by dipole monolayers on the gate insulator," Journal of Applied Physics, vol. 96, pp.

6431-6438, 2004.

[64] S. Kobayashi, et al., "Control of carrier density by self-assembled monolayers in

organic field-effect transistors," Nature Materials, vol. 3, pp. 317-322, 2004.

[65] A. Salleo, et al., "Polymer thin-film transistors with chemically modified dielectric

interfaces," Applied Physics Letters, vol. 81, pp. 4383-4385, 2002.

[66] Z. Liu, et al., "Performance and Reliability of Organic Transistors Controlled by

Interfacial Phenyl-Terminated Self-Assembled Monolayers," Materials Research

Society (MRS) Fall Meeting, Boston, MA, Dec. 1-5 2008.

[67] Z. Liu, et al., Manuscript in preparation, 2009.

[68] D. Kumaki, et al., "Influence of H2O and O2 on threshold voltage shift in

organic thin-film transistors: Deprotonation of SiOH on SiO2 gate-insulator

surface," Applied Physics Letters, vol. 92, p. 093309, Mar 2008.

[69] J. Peanasky, et al., "Self-Assembled Monolayers on Mica for Experiments

Utilizing the Surface Forces Apparatus," Langmuir, vol. 11, pp. 953-962, 1995.

132 BIBLIOGRAPHY

[70] H. Y. Nie, et al., "Delivering octadecylphosphonic acid self-assembled

monolayers on a Si wafer and other oxide surfaces," Journal of Physical Chemistry B,

vol. 110, pp. 21101-21108, Oct 2006.

[71] Y. Ito, et al., "Crystalline Ultrasmooth Self-Assembled Monolayers of

Alkylsilanes for Organic Field-Effect Transistors," Journal of the American Chemical

Society, vol. 131, pp. 9396-9404, Jul 2009.

[72] Z. Liu, et al., Manuscript in preparation, 2009.

[73] I. Mcculloch, et al., "Liquid-crystalline semiconducting polymers with high

charge-carrier mobility," Nature Materials, vol. 5, pp. 328-333, Apr 2006.

[74] C. Krzeminski, et al., "Theory of electrical rectification in a molecular

monolayer," Physical Review B, vol. 6408, pp. -, Aug 15 2001.

[75] Z. Liu, et al., Manuscript in preparation, 2009.

[76] Y. Y. Noh, et al., "Downscaling of self-aligned, all-printed polymer thin-film

transistors," Nature Nanotechnology, vol. 2, pp. 784-789, Dec 2007.

[77] T. Minari, et al., "Scaling effect on the operation stability of short-channel

organic single-crystal transistors," Applied Physics Letters, vol. 91, p. 063506, Aug

2007.

[78] G. S. Tulevski, et al., "Device scaling in sub-100 nm pentacene field-effect

transistors," Applied Physics Letters, vol. 89, p. 183101, Oct 2006.

[79] J. B. Lee, et al., "10-nm channel length pentacene transistors," IEEE Transactions

on Electron Devices, vol. 52, pp. 1874-1879, Aug 2005.

[80] L. Wang, et al., "Pentacene field-effect transistors with sub-10-nm channel

lengths," Applied Physics Letters, vol. 85, pp. 1772-1774, Sep 2004.

[81] M. Leufgen, et al., "Optimized sub-micron organic thin-film transistors: the

influence of contacts and oxide thickness," Synthetic Metals, vol. 146, pp. 341-345,

Nov 2004.

BIBLIOGRAPHY 133

[82] J. Zaumseil, et al., "Nanoscale organic transistors that use source/drain

electrodes supported by high resolution rubber stamps," Applied Physics Letters,

vol. 82, pp. 793-795, Feb 2003.

[83] J. Collet, et al., "Low-voltage, 30 nm channel length, organic transistors with a

self-assembled monolayer as gate insulating films," Applied Physics Letters, vol. 76,

pp. 1941-1943, Apr 2000.

[84] J. A. Rogers, et al., "Low-voltage 0.1 mu m organic transistors and

complementary inverter circuits fabricated with a low-cost form of near-field

photolithography," Applied Physics Letters, vol. 75, pp. 1010-1012, Aug 1999.

[85] X. H. Zhang and B. Kippelen, "High-performance C-60 n-channel organic field-

effect transistors through optimization of interfaces," Journal of Applied Physics,

vol. 104, p. 104504, Nov 2008.

[86] E. C. P. Smits, et al., "Bottom-up organic integrated circuits," Nature, vol. 455, pp.

956-959, Oct 2008.

[87] T. Maeda, et al., "Organic field-effect transistors with reduced contact

resistance," Applied Physics Letters, vol. 89, p. 123508, Sep 2006.

[88] D. J. Gundlach, et al., "An experimental study of contact effects in organic thin

film transistors," Journal of Applied Physics, vol. 100, p. 024509, Jul 2006.

[89] E. J. Meijer, et al., "Scaling behavior and parasitic series resistance in disordered

organic field-effect transistors," Applied Physics Letters, vol. 82, pp. 4576-4578, Jun

2003.

[90] H. Klauk, et al., "Contact resistance in organic thin film transistors," Solid-State

Electronics, vol. 47, pp. 297-301, Feb 2003.

[91] B. H. Hamadani, et al., "Undoped polythiophene field-effect transistors with

mobility of 1 cm(2) V-1 s(-1)," Applied Physics Letters, vol. 91, p. 243512, Dec

2007.

[92] S. Verlaak, et al., "Modeling of transport in polycrystalline organic semiconductor

films," Applied Physics Letters, vol. 82, pp. 745-747, Feb 2003.

134 BIBLIOGRAPHY

[93] S. Scheinert and G. Paasch, "Interdependence of contact properties and field-

and density-dependent mobility in organic field-effect transistors," Journal of

Applied Physics, vol. 105, p. 014509, Jan 2009.

[94] P. V. Necliudov, et al., "Modeling of organic thin film transistors of different

designs," Journal of Applied Physics, vol. 88, pp. 6594-6597, Dec 2000.

[95] I. Kymissis, et al., "High-performance bottom electrode organic thin-film

transistors," IEEE Transactions on Electron Devices, vol. 48, pp. 1060-1064, Jun

2001.

[96] T. Li, et al., "Investigation of bottom-contact organic field effect transistors by

two-dimensional device modeling," Journal of Applied Physics, vol. 93, pp. 4017-

4022, Apr 2003.

[97] M. Koehler, et al., "Resolving the contact voltage dilemma in organic field effect

transistors," Physical Review B, vol. 78, p. 153312, Oct 2008.

[98] J. H. Werner and H. H. Guttler, "Barrier Inhomogeneities at Schottky Contacts,"

Journal of Applied Physics, vol. 69, pp. 1522-1533, Feb 1991.

[99] Y. N. Gartstein and E. M. Conwell, "Field-dependent thermal injection into a

disordered molecular insulator," Chemical Physics Letters, vol. 255, pp. 93-98, Jun

1996.

[100] V. I. Arkhipov, et al., "Current injection from a metal to a disordered hopping

system. II. Comparison between analytic theory and simulation," Physical Review B,

vol. 59, pp. 7514-7520, Mar 1999.

[101] V. I. Arkhipov, et al., "Charge injection into light-emitting diodes: Theory and

experiment," Journal of Applied Physics, vol. 84, pp. 848-856, Jul 1998.

[102] S. Barth, et al., "Current injection from a metal to a disordered hopping system.

III. Comparison between experiment and Monte Carlo simulation," Physical

Review B, vol. 60, pp. 8791-8797, Sep 1999.

BIBLIOGRAPHY 135

[103] U. Wolf, et al., "Current injection from a metal to a disordered hopping system. I.

Monte Carlo simulation," Physical Review B, vol. 59, pp. 7507-7513, Mar 1999.

[104] M. A. Baldo and S. R. Forrest, "Interface-limited injection in amorphous organic

semiconductors," Physical Review B, vol. 6408, p. 085201, Aug 2001.

[105] T. P. Nguyen and J. L. Mansot, "Metal-oxygen-carbon interaction in the poly(p-

phenylene vinylene)-aluminum system: A study by analytical transmission

electron microscopy and X-ray photoelectron spectroscopy," Thin Solid Films, vol.

283, pp. 135-139, Sep 1996.

[106] M. Koehler, et al., "Electrical characteristics in unipolar conjugated polymer

devices: the case of modified transport properties in the neighbourhood of the

top electrode/polymer interface," Journal of Physics D-Applied Physics, vol. 34, pp.

1947-1950, Jul 2001.

[107] T. Li, et al., "Channel formation in organic field-effect transistors," Journal of

Applied Physics, vol. 91, pp. 4312-4318, Apr 2002.

[108] P. L. Bullejos, et al., "Model for the injection of charge through the contacts of

organic transistors," Journal of Applied Physics, vol. 105, p. 084516, Apr 2009.

[109] B. H. Hamadani and D. Natelson, "Nonlinear charge injection in organic field-

effect transistors," Journal of Applied Physics, vol. 97, p. 064508, Mar 2005.

[110] B. H. Hamadani and D. Natelson, "Extracting contact effects in organic FETs,"

Proceedings of the IEEE, vol. 93, pp. 1306-1311, Jul 2005.

[111] B. H. Hamadani and D. Natelson, "Temperature-dependent contact resistances

in high-quality polymer field-effect transistors," Applied Physics Letters, vol. 84, pp.

443-445, Jan 2004.

[112] L. Bürgi, et al., "Close look at charge carrier injection in polymer field-effect

transistors," Journal of Applied Physics, vol. 94, pp. 6129-6137, Nov 2003.

[113] R. A. Street and A. Salleo, "Contact effects in polymer transistors," Applied

Physics Letters, vol. 81, pp. 2887-2889, Oct 2002.

136 BIBLIOGRAPHY

[114] K. Seshadri and C. D. Frisbie, "Potentiometry of an operating organic

semiconductor field-effect transistor," Applied Physics Letters, vol. 78, pp. 993-995,

Feb 2001.

[115] A. B. Chwang and C. D. Frisbie, "Field effect transport measurements on single

grains of sexithiophene: Role of the contacts," Journal of Physical Chemistry B, vol.

104, pp. 12202-12209, Dec 2000.

[116] J. C. Scott, "Metal-organic interface and charge injection in organic electronic

devices," Journal of Vacuum Science & Technology A, vol. 21, pp. 521-531, May-Jun

2003.

[117] S. M. Sze and K. K. Ng, Physics of Semiconductor Devices, 3rd ed. Hoboken, NJ:

Wiley-Interscience, 2007.

[118] J. G. Simmons, "Richardson-Schottky Effect in Solids," Physical Review Letters, vol.

15, pp. 967-968, 1965.

[119] P. R. Emtage and J. J. O'Dwyer, "Richardson-Schottky Effect in Insulators,"

Physical Review Letters, vol. 16, pp. 356-358, 1966.

[120] C. R. Crowell and S. M. Sze, "Current Transport in Metal-Semiconductor

Barriers," Solid-State Electronics, vol. 9, pp. 1035-1048, 1966.

[121] J. C. Scott and G. G. Malliaras, "Charge injection and recombination at the

metal-organic interface," Chemical Physics Letters, vol. 299, pp. 115-119, Jan 1999.

[122] Y. L. Shen, et al., "Mobility-dependent charge injection into an organic

semiconductor," Physical Review Letters, vol. 86, pp. 3867-3870, Apr 2001.

[123] M. L. Chabinyc, et al., "Connecting electrical and molecular properties of

semiconducting polymers for thin-film transistors," MRS Bulletin, vol. 33, pp.

683-689, Jul 2008.

[124] V. I. Arkhipov, et al., "Effect of doping on the density-of-states distribution and

carrier hopping in disordered organic semiconductors," Physical Review B, vol. 71,

p. 045214, Jan 2005.

BIBLIOGRAPHY 137

[125] W. Warta and N. Karl, "Hot Holes in Naphthalene - High, Electric-Field-

Dependent Mobilities," Physical Review B, vol. 32, pp. 1172-1182, 1985.

[126] A. L. Briseno, et al., "Patterning organic single-crystal transistor arrays," Nature,

vol. 444, pp. 913-917, Dec 2006.

[127] V. C. Sundar, et al., "Elastomeric transistor stamps: Reversible probing of charge

transport in organic crystals," Science, vol. 303, pp. 1644-1646, Mar 2004.

[128] T. van Woudenbergh, et al., "Temperature dependence of the charge injection in

poly-dialkoxy-p-phenylene vinylene," Applied Physics Letters, vol. 79, pp. 1697-

1699, Sep 2001.

[129] P. W. Anderson, "Absence of Diffusion in Certain Random Lattices," Physical

Review, vol. 109, pp. 1492-1505, 1958.

[130] A. Miller and E. Abrahams, "Impurity Conduction at Low Concentrations,"

Physical Review, vol. 120, pp. 745-755, 1960.

[131] N. F. Mott and E. A. Davis, Electronic Processes in Non-Crystalline Materials. Oxford:

Clarendon Press, 1971.

[132] M. C. J. M. Vissenberg and M. Matters, "Theory of the field-effect mobility in

amorphous organic transistors," Physical Review B, vol. 57, pp. 12964-12967, May

1998.

[133] A. Salleo, et al., "Intrinsic hole mobility and trapping in a regioregular

poly(thiophene)," Physical Review B, vol. 70, p. 115311, Sep 2004.

[134] R. A. Street, Hydrogenated amorphous silicon. Cambridge ; New York: Cambridge

University Press, 1991.

[135] P. G. Lecomber and W. E. Spear, "Electronic Transport in Amorphous Silicon

Films," Physical Review Letters, vol. 25, pp. 509-511, 1970.

[136] G. Horowitz, et al., "Temperature and gate voltage dependence of hole mobility

in polycrystalline oligothiophene thin film transistors," Journal of Applied Physics,

vol. 87, pp. 4456-4463, May 2000.

138 BIBLIOGRAPHY

[137] J. F. Chang, et al., "Relative importance of polaron activation and disorder on

charge transport in high-mobility conjugated polymer field-effect transistors,"

Physical Review B, vol. 76, p. 205204, Nov 2007.

[138] R. A. Street, et al., "Transport in polycrystalline polymer thin-film transistors,"

Physical Review B, vol. 71, p. 165202, Apr 2005.

[139] H. Bässler, "Charge Transport in Disordered Organic Photoconductors - a

Monte-Carlo Simulation Study," Physica Status Solidi B-Basic Research, vol. 175, pp.

15-56, Jan 1993.

[140] S. J. Martin, et al., "Temperature and field dependence of the mobility of highly

ordered conjugated polymer films," Physical Review B, vol. 67, p. 165214, Apr

2003.

[141] S. V. Novikov, et al., "Essential role of correlations in governing charge transport

in disordered organic materials," Physical Review Letters, vol. 81, pp. 4472-4475,

Nov 1998.

[142] R. E. Parris, et al., "Nature of charge carriers in disordered molecular solids: Are

polarons compatible with observations?," Physical Review Letters, vol. 8712, p.

126601, Sep 2001.

[143] A. J. Mozer, et al., "Charge carrier mobility in regioregular poly(3-hexylthiophene)

probed by transient conductivity techniques: A comparative study," Physical

Review B, vol. 71, p. 035214, Jan 2005.

[144] T. Kreouzis, et al., "Temperature and field dependence of hole mobility in

poly(9,9-dioctylfluorene)," Physical Review B, vol. 73, p. 235201, Jun 2006.

[145] P. W. M. Blom, et al., "Electric-field and temperature dependence of the hole

mobility in poly(p-phenylene vinylene)," Physical Review B, vol. 55, pp. R656-R659,

Jan 1997.

BIBLIOGRAPHY 139

[146] W. D. Gill, "Drift Mobilities in Amorphous Charge-Transfer Complexes of

Trinitrofluorenone and Poly-N-Vinylcarbazole," Journal of Applied Physics, vol. 43,

pp. 5033-5040, 1972.

[147] H. Klauk (Eds.), Organic Electronics: Materials, Manufacturing and Applications:

Weinheim, Germany: Wiley-VCH Verlag GmbH, 2006.

[148] T. Hassenkam, et al., "Direct Visualization of the Nanoscale Morphology of

Conducting Polythiophene Monolayers Studied by Electrostatic Force

Microscopy," Advanced Materials, vol. 13, pp. 631-634, 2001.

[149] L. Burgi, et al., "Noncontact potentiometry of polymer field-effect transistors,"

Applied Physics Letters, vol. 80, pp. 2913-2915, Apr 2002.

[150] Y. He, et al., "Schottky barrier formation at metal electrodes and semiconducting

carbon nanotubes," Applied Physics Letters, vol. 94, p. 093107, Mar 2009.

[151] X. Guan, et al., "Modeling of schottky and ohmic contacts between metal and

graphene nanoribbons using extended h&#x00FC;ckel theory (EHT)-based

NEGF method," in Technical Digest of IEEE International Electron Devices Meeting

(IEDM), 2008, pp. 197-200.

[152] S. Picozzi, et al., "Schottky barrier height at an organic/metal junction: A first-

principles study of PTCDA/X (X = Al, Ag) contacts," Physical Review B, vol. 68,

p. 195309, Nov 2003.

[153] I. G. Hill, et al., "Molecular level alignment at organic semiconductor-metal

interfaces," Applied Physics Letters, vol. 73, pp. 662-664, Aug 1998.

[154] H. Ishii, et al., "Energy level alignment and interfacial electronic structures at

organic/metal and organic/organic interfaces (vol 11, pg 605, 1999)," Advanced

Materials, vol. 11, pp. 972-972, Aug 1999.

[155] S. Braun, et al., "Energy-Level Alignment at Organic/Metal and

Organic/Organic Interfaces," Advanced Materials, vol. 21, pp. 1450-1472, Apr

2009.

140 BIBLIOGRAPHY

[156] C. S. Lee, et al., "Interface dipole at metal-organic interfaces: Contribution of

metal induced interface states," Applied Physics Letters, vol. 94, p. 113304, Mar

2009.

[157] V. Heine, "Theory of Surface States," Physical Review, vol. 138, pp. A1689-A1696,

1965.

[158] D. Connelly, et al., "New route to zero-barrier metal source/drain MOSFETs,"

IEEE Transactions on Nanotechnology, vol. 3, pp. 98-104, Mar 2004.

[159] M. Kobayashi, et al., "Fermi-level depinning in metal/Ge Schottky junction and

its application to metal source/drain Ge NMOSFET," in IEEE Symposium on

VLSI Technology (VLSI), 2008, pp. 54-55.