You are on page 1of 172

TFT/LCD

Japanese Technology Reviews


Editor in Chief

Toshiaki Ikoma, University of Tokyo

Section Editors

Section A: Electronics

Toshiaki Ikoma, University of Tokyo

Section B: Computers and Communications

Tadao Saito, University of Tokyo

Kazumoto Iinuma, NEC Corporation, Kawasaki

Section C: New Materials

Hiroaki Yanagida, University of Tokyo

Noboru Ichinosé, Waseda University, Tokyo

Section D: Manufacturing Engineering

Fumio Harashima, University of Tokyo

Section E: Biotechnology

Isao Karube, University of Tokyo

Reiko Kuroda, University of Tokyo

Section A: Electronics

Volume 2
MMIC-Monolithic Microwave Integrated Circuits
Yasuo Mitsui

Volume 4
Bulk Crystal Growth Technology
Shin-ichi Akai, Keiichiro Fujita, Masamichi Yokogawa, Mikio Morioko and Kazuhisa Matsumoto

Volume 8
Semiconductor Heterostructure Devices
Masayuki Abe and Naoki Yokoyama

Volume 11
Development of Optical Fibers in Japan
Hiroshi Murata

Volume 12
High-Performance BiCMOS Technology and Its Applications to VLSIs
Ikuro Masuda and Hideo Maejima

Volume 13
Semiconductor Devices for Electronic Tuners
Seiichi Watanabe

Volume 19
Amorphous Silicon Solar Cells
Yukinori Kuwano

Volume 20
High Density Magnetic Recording for Home VTR: Heads and Media
Kazunori Ozawa

Volume 29
TFT/LCD
Toshihisa Tsukada

TFT/LCD

Liquid-Crystal Displays Addressed by Thin-Film Transistors

Toshihisa Tsukada

Hitachi Ltd.
Tokyo, Japan

Gordon and Breach Science Publishers

Australia • Canada • France • Germany • India • Japan • Luxembourg


Malaysia • The Netherlands • Russia • Singapore • Switzerland

Copyright © 1996 OPA (Overseas Publishers Association) N.V.


Published by license under the Gordon and Breach Science Publishers
imprint.

All rights reserved.


First published 1996

This edition published in the Taylor & Francis e-Library, 2003.

Second printing 2000

No part of this book may be reproduced or utilized in any form or by any means, electronic or mechanical, including
photocopying and recording, or by any information storage or retrieval system, without permission in writing from the
publisher. Printed in Malaysia.

Amsteldijk 166

1st Floor

1079 LH Amsterdam

The Netherlands

British Library Cataloguing in Publication Data


Tsukada, Toshihisa
TFT/LCD: Liquid-Crystal Displays
Addressed by Thin-film Transistors.—
(Japanese Technology Reviews. v. 29)
I. Title II. Series
621.38152

ISBN 0-203-41326-1 Master e-book ISBN

ISBN 0-203-41331-8 (MP PDA Format)

ISBN 2-919875-01-9 (Print Edition)

Copyright © 2002/2003 Mobipocket.com. All rights reserved.

Reader's Guide

This ebook has been optimized for MobiPocket PDA.

Tables may have been presented to accommodate this Device's Limitations.

Table content may have been removed due to this Device's Limitations.

Image presentation is limited by this Device's Screen resolution.

All possible language characters have been included within the Font handling ability of this Device.

Contents
Preface to the Series

Preface
1. Introduction

2. TFT/LCD

2.1. TFT/LCD Configuration

2.2. Pixel Design

2.3. Design Analysis

2.3.1. Charging of pixel capacitance

2.3.2. Voltage offset

2.3.3. Gate delay

2.3.4. Colorimetric design

2.4. TFT Panel Fabrication

References

3. Thin-Film Transistors

3.1. Hydrogenated Amorphous Silicon Thin-Film Transistors

3.2. TFT Characteristics

3.2.1. Current-voltage characteristics

3.2.2. Parasitic effects

3.3. Threshold Voltage Shift

3.4. Process-Related Issues

3.4.1. Aluminum gate TFT

3.4.2. Self-alignment process

3.5. Computer Simulation of a-Si:H TFT

References

4. Hydrogenated Amorphous Silicon

4.1. Amorphous Semiconductors

4.2. Density of States of a-Si:H

4.3. Conductivity and Carrier Mobility

4.4. Absorption Spectra of a-Si:H

4.4.1. Optical absorption

4.4.2. IR absorption
References

5. Twisted-Nematic Cell

5.1. Threshold Voltage of TN Cell

5.2. C-V Characteristics of TN Cell

5.3. Optical Properties of TN Cell

5.4. Super-Twisted Nematic (STN) Cell

References

6. Liquid Crystal

6.1. Dielectric Permittivity and Refractive Index

6.2. Elastic Constants

6.3. Viscosity

References

General References

Index

Preface to the Series


Modern technology has a great impact on both industry and society. New technology is first created by pioneering
work in science. Eventually, a major industry is born, and it grows to have an impact on society in general.
International cooperation in science and technology is necessary and desirable as a matter of public policy. As
development progresses, international cooperation changes to international competition, and competition further
accelerates technological progress.

Japan is in a very competitive position relative to other developed countries in many high-technology fields. In some
fields, Japan is in a leading position: for example, manufacturing technology and micro-electronics, especially
semiconductor LSIs and optoelectronic devices. Japanese industries lead in the application of new materials such as
composites and fine ceramics, although many of these new materials were first developed in the United States and
Europe. The United States, Europe and Japan are working intensively, both competitively and cooperatively, on the
research and development of high-critical-temperature superconductors. Computers and communications are now a
combined field that plays a key role in the present and future of human society. In the next century, biotechnology will
grow, and it may become a major segment of industry. While Japan does not play a major role in all areas of
biotechnology, in some areas such as fermentation (the traditional technology for making sake), Japanese research is of
primary importance.

Today, tracking Japanese progress in high-technology areas is both a necessary and rewarding process. Japanese
academic institutions are very active; consequently, their results are published in scientific and technical journals and
are presented at numerous meetings where more than 20,000 technical papers are presented orally every year.
However, due principally to the language barrier, the results of academic research in Japan are not well-known
overseas. Many in the United States and in Europe are thus surprised by the sudden appearance of Japanese high-
technology products. The products are admired and enjoyed, but some are astonished at how suddenly these products
appear.

With the series Japanese Technology Reviews, we present state-of-the-art Japanese technology in five fields:

Electronics

Computers and Communications

New Materials

Manufacturing Engineering

Biotechnology

Each tract deals with one topic within each of these five fields and reviews both the present status and future prospects
of the technology, mainly as seen from the Japanese perspective. Each author is an outstanding scientist or engineer
actively engaged in relevant research and development.

The editors are confident that this series will not only give a deep insight into Japanese technology but will also be
useful for developing new technology of interest to our readers.

As editor-in-chief, I would like to sincerely thank the members of the editorial board and the authors for their
contributions to this series.

TOSHIAKI IKOMA

Preface
This is the first book describing in full detail liquid-crystal displays addressed by thin-film transistors (TFT/LCDs):
their design, fabrication, characteristics, material issues, as well as their basic operation. Since the fields relevant to
TFT/LCDs are quite diverse, the topics covered in this book are interdisciplinary. The two most important materials
discussed are amorphous semiconductors (a-Si:H) and liquid crystals. The unique features of TFT/LCDs arise from the
combination of these two materials. In this book, device characteristics and material issues are described in terms of
how they relate to display technology. Therefore, themes not directly related to TFT/LCDs but quite important in other
applications were left to the other excellent books listed below. In conventional books, the opening chapters deal with
basic studies and later chapters emphasize application. However, this book starts with an overview of the TFT/LCD,
including its design, fabrication and characteristics; then, the following chapters describe devices, materials and their
basic principles. This arrangement will make it easier for the reader to fully understand the contents of this book by
first explaining the overall subject matter.

This book will be useful to those who are engaged in the research and development of TFT/LCDs. It will be of great
interest especially to those who are working in the fields of consumer electronics and computer and communication
engineering, since the display quality of the man-machine interface is critical in these applications. Those who are
studying solid-state physics, mesophase chemistry, and electronic devices and systems, will also find information of
great interest.

I would like to express my gratitude for the support that I have received from Hitachi Ltd. and from my many
colleagues at the Central Research Laboratory, Hitachi Research Laboratory, and Mobara Works of Hitachi Ltd.
Particular thanks go to E.Kawamata who helped to prepare the manuscript of this book.

T.Tsukada

Kokubunji 1994
CHAPTER 1
Introduction
The TFT/LCD, an abbreviation of thin-film-transistor-addressed liquid-crystal display, is a flat-panel display in which
the display medium is liquid-crystal and each picture element (pixel) is controlled by a thin-film transistor. The
TFT/LCD is creating a whole new world of technology in consumer electronics and in computer and communication
systems. The market for TFT/LCDs is now growing much faster than expected and impacting new application fields,
as well as conventional fields.

The concept of the TFT/LCD is not new, but rather old. As early as 1966, Weimer 1 mentioned the possibility of using
TFTs as display switches. A more detailed concept was described by Lechner et al. 2 in 1971, where the use of diodes
or triodes (transistors) was discussed as switches for active-matrix liquid-crystal displays. The use of storage
capacitors implemented in parallel with the liquid-crystal cell capacitor was also mentioned. Preceding this discussion,
Heilmeier et al. 3 proposed nematic liquid-crystal as a material for a flat-panel display. A sandwich cell consisting of a
transparent front electrode, a reflecting back electrode, and nematic liquid-crystal in between, was prepared (reflective
mode). When there was no applied field, the cell appeared black. When a dc voltage was applied, the liquid-crystal
became turbulent and scattered light: the cell appeared white. This phenomenon was termed "dynamic scattering", and
was applied to demonstrate the first liquid-crystal display both in reflective and transmissive mode operations. A
3.5"×4" alphanumeric display was fabricated and a maximum contrast ratio higher than 20 to 1, was demonstrated.

Brody et al. 4 applied the CdSe TFT to the active-matrix liquid-crystal panel. This display panel consisted of 14,000
transistors, storage capacitors and the twisted-nematic (TN) liquid-crystal cell. Although TFTs were made of CdSe
rather than a-Si:H, the configuration is essentially the same as today's TFT/LCD panels. The TN liquid-crystal cell was
first proposed by Schadt and Helfrich 5 and featured low-voltage operation, low power consumption, and fast response
time. In the TN cell, the average direction (director) of the liquid-crystal molecules is twisted 90° as they go from the
back to the front glass substrate. The polarization of the light is also rotated 90° as the light passes though the liquid-
crystal cell. When the front polarizer is set parallel to the rear polarizer, the light is not transmitted as there is no
applied voltage. When the ac voltage is applied to the cell, the director of liquid-crystal molecules becomes
perpendicular to the substrate. In this state, the TN cell becomes optically inactive and linearly polarized light travels
through the cell without any rotation of polarization. Therefore, the light is transmitted through the front polarizer.
Since its development, the TN cell has played a very important role as the display medium for TFT/LCDs.

Hydrogenated amorphous silicon (a-Si:H) was a late arrival in TFT technologies. However, it had a great influence in
achieving practical TFT/LCDs. Since the first report by the Dundee group 6 , a-Si:H TFT has been recognized as the
most suitable device for TFT/LCDs. The mobility of a-Si:H has proved to be just enough to charge liquid-crystal
capacitance and storage capacitor. The off-current of a-Si:H TFT has turned out to be on a sufficiently low level not to
discharge the pixel capacitance during the frame time, owing to the high resistivity of undoped a-Si:H. Another
important feature of a-Si:H is that it can be deposited over a large area. Uniformity over the area, and reproducibility
from run to run can be obtained relatively easily. Also important is that it can be deposited at low temperatures, so
glass substrates can be used. Moreover, good interface properties between a-Si:H and other thin films, like metals,
insulators, and semiconductors can be obtained. Due to these properties, the hydrogenated amorphous silicon TFT has
acquired wide acceptance among the many candidates to make TFT/LCDs.

As for liquid-crystal, more than one hundred years have passed since its discovery. In 1888, an Austrian botanist, F.
Reinitzer reported that cholesteric benzoate which he had purified, showed a strong birefringence at a temperature
range between 145.5°C and 178.5°C. He sent the sample, which became a turbulent liquid in this temperature range, to
a German physicist O.Lehmann. Lehmann found that this material showed an optical anisotropy when observed under
a crossed Nikol and coined the term "liquid-crystal" for the first time. It took a long (eighty-year) elapse of time after
this discovery until the concept of liquid-crystals appeared as a medium for displays. Even after this, about twenty
years were necessary to develop the practical TFT/LCD. However, we are now at the stage to benefit from this long
history of research and development, which will be described in the following chapters.
References
1. Weimer, P.K. (1966). Thin film transistors. In Field Effect Transistors , edited by J.T.Wallmark and H.Johnson.
New Jersey: Prentice Hall.

2. Lechner, B.J., Marlowe, F.J., Nester, E.O., and Tults, J. (1971). liquid-crystal matrix displays. Proc. IEEE , 59,
1566–1579.

3. Heilmeier, G.H., Zanoni, L.A., and Barton, L.A. (1968). Dynamic scattering: a new electrooptic effect in certain
classes of nematic liquid-crystals. Proc. IEEE , 56, 1162–1171.

4. Brody, T.P., Asars, J.A., and Dixon, G.D. (1973). A 6×6 inch 20 lines-per-inch liquid-crystal display panel. IEEE
Trans. Electron Devices , ED-20, 995–1001.

5. Schadt, M. and Helfrich, W. (1971). Voltage-dependent optical activity of a twisted nematic liquid-crystal. Applied
Physics Letters , 18, 127–128.

6. LeComber, P.G., Spear, W.E., and Gaith, A. (1979). Amorphous-silicon field-effect device and possible application.
Electronics Letters , 15, 179–181.

CHAPTER 2
TFT/LCD
The thin-film-transistor-addressed liquid-crystal display (TFT/LCD) is a flat-panel display which is used in a variety
of products, including consumer electronics, computers, and communication terminals. However, the application of
TFT/LCD to such products is only a recent development. For more than forty years, cathode-ray tubes (CRTs) have
dominated the display world. The CRT's most dramatic success has been in monochrome and then color TVs. With the
expansion of the computer industry, computer terminal displays have also provided a large market for CRTs. As a
result, the word "display" has become almost synonymous with "CRT".

However, the situation is now changing rapidly. Developments in transistors and integrated circuits have made it
possible to greatly reduce the size of computers due to improvements in the performance of miroprocessors,
semiconductor memories, and other devices. As a result, main-frame computers have given way to minicomputers, to
workstations, and to personal computers. This evolution hastened the arrival of flat-panel displays, including
TFT/LCDs. Not only do these displays avoid the curvature normally associated with CRTs, they are also lightweight
and consume little power. These features make them ideal for use in modern electronic systems.

TFT/LCDs are generally characterized by the diagonal length of the panel and their resolution. Figure 2.1 shows the
correspondence between display area and the number of pixels (picture elements). The numerical examples shown in
this figure correspond to actual video-data terminals. In computer display applications each pixel generally consists of
three colored stripes (one red, one green, and one blue). The inserts in this figure show the relative sizes of this
configuration, assuming the pixels are square. A resolution of 640 (horizontal) by 480 (vertical) pixels corresponds to a
video-graphic-array (VGA) monitor, and if a pixel size of 0.33 mm square is assumed, diagonal of VGA monitors
becomes 26.4 cm or 10.4". In panels with a resolution of
Figure 2.1 Number of pixels vs. area of flat-panel displays. The number of pixels in rows and columns shown in this
figure by H and V correspond to computer-terminal-display applications. At each resolution, an example is shown for
the panel and pixel size for an assumed substrate size of 460×360 mm2 .

1,280 (H) by 1,024 (V), the number of pixels exceeds one million, and there are more than three million RGB dots or
subpixels. Thus, more than three million TFTs must be fabricated on these panels.

2.1. TFT/LCD Configuration


The basic configuration of a TFT/LCD is shown in Figure 2.2. Liquid-crystal is encapsulated between two glass
substrates, a TFT substrate and a color-filter substrate. The color-filter substrate is also called the common electrode
substrate. The transparent common electrode on this substrate is made of ITO (Indium Tin Oxide), and is deposited on
top of the color filter. In order to obtain good display quality, the cell gap of the liquid-crystal (i.e., the spacing
between the two glass substrates) has to be precisely controlled to a specific value, e.g., 5 um. This gap has to be
uniform over the whole display area and reproducible from
Figure 2.2 Configuration of TFT/LCD. The TFT and the color-filter substrates are two parallel sheets of glass with
liquid crystal injected between them. A crossed-polarizer system is shown here, corresponding to a normally-white
display.

run to run. Therefore, transparent spacers such as plastic beads are placed on the surface of the glass substrate.

The liquid-crystal cells are twisted-nematic type (see Chapter 5) in which the director (orientation) of the liquid-
crystal molecules is twisted 90° between the TFT substrate and the common electrode substrate. In Figure 2.2, the
crossed-polarizer system is shown, in which the first polarizer works as a backlight polarizer and the other acts as an
analyzer. In this system, light passes through the analyzer when there is no applied voltage on the cell, and is blocked
when the applied voltage is high enough to align the liquid-crystal molecules vertically. The liquid-crystal is anchored
on the surface of the glass substrates so that its molecules are oriented to a proper direction. In order to set the
anchoring direction, the glass substrate is coated with an organic film such as a polyimide film and the surface of the
film is rubbed with a fabric in a specific direction. The liquid-crystal molecules are tilted several degrees with respect
to the glass surface. This tilt angle is called the pretilt angle and plays an important role in determining the electrical
Figure 2.3 Schematic diagram of the display system. Controllers, a power supply, and other circuitry are combined to
operate the display.

and optical characteristics of the TFT/LCD, which will be described in the following sections.

The TFT substrate consists of a TFT array and an array of external terminals on which LSIs are bonded to drive the
TFT panel. The driver LSIs are essentially scan generators for the horizontal and vertical buslines. These LSIs are
directly bonded to the glass with TAB (Tape-Automated Bonding) connectors, and they provide each pixel of the
panel with video signals that are transferred to the panel via a video
Figure 2.4 The brightness of the display module is much lower than that of the backlight illumination. Only 5% or so
of the original brightness is output from the front polarizer.

signal processor and controller. A schematic diagram of TFT/LCD module and controllers is shown in Figure 2.3.

The backlight system can be either direct or indirect. With direct lighting, one or more fluorescent lamps are
positioned directly beneath the rear polarizer, and with indirect lighting, a light-guide is used to guide the light from
lamp(s) situated beside it. The backlight illumination is attenuated as it passes through the display module as shown in
Figure 2.4. The maximum transmittances of the polarizer and color filter are one-half and one-third, respectively,
resulting in a utility factor of one-sixth. The aperture ratio of the pixels further reduces this factor. If an aperture ratio
of 50% is assumed, the utility factor will be 8%. However, this is only the upper limit of the transmittance of the total
system. In a practical system, the total utility factor is 3–6%. An example of a backlight spectrum is shown in Figure
2.5 1 . Three-wave-length-type illumination is generally used as the backlight.

Figure 2.6 shows a schematic diagram of a TFT/LCD. There are two sets of buslines, i.e., horizontal gate buslines and
vertical data buslines.
Figure 2.5 An example of a backlight illumination spectrum. The three wave-lengths of R, G, and B are designed so
as to obtain good color quality of the display panel.

Figure 2.6 Schematic diagram of a TFT/LCD. In this particular configuration for a computer terminal display, each
pixel is designed to be square-shaped and consists of three RGB subpixels or dots.

A TFT is formed at each intersection of these buslines to turn on and off the voltage applied to the liquid-crystal cell.
This cell is represented by an equivalent capacitance and in parallel with this capacitor a storage capacitor or additional
capacitor (Cst or Cadd) is formed to improve the retention characteristics of the signal charge. Details of this capacitor
will be given in the following sections. The color filter is formed in a striped R, G and B configuration, three dots
forming one pixel. The display operates one line at a time—video signals are fed to the data buslines simultaneously
through a data buffer during the gate turn-on time. The scan gate voltage pulse applied to a certain (say, i-th) gate

Figure 2.7 Timing chart of the TFT/LCD. The panel is refreshed at a rate of 60–70 frames per second. For a display
with 480 gate buslines and 60 frames per second, the gate select time or the gate pulse width is 34.7 us. The pixel
electrodes must be fully charged to the signal voltage during this period, and must also keep this signal voltage until
the next charging step arrives.

busline opens the gates of the TFTs connected to this busline. The signal voltage is then applied to the pixel electrode
of each dot on this gate busline. The timing chart for scanning the TFT/LCD is shown in Figure 2.7. The period t ON to
turn on the TFTs on the i-th gateline, is given by

(2.1)

where m is the number of gatelines and f F is the frame frequency. If the frame frequency is 60 Hz and if there are 480
gate buslines, t ON is 34.7 us. During this time period, the charging of the capacitance (the liquid-crystal cell and the
storage capacitor) has to be completed.

After this charging period, the liquid-crystal cells on the i-th gate line are cut off from the data lines and the cells
connected to the (i+l)-th gate lines are charged. The cell cutoff has to be perfect, i.e., the cutoff cell has to keep its
charged voltage until the next charging step takes place. If, for some reason, there is an increase of the off-current of
the TFT, the signal voltage will discharge causing crosstalk and degraded display quality. In fact, the most frequent
cause of crosstalk is an increase in leakage current due to the photocurrent induced by intense illumination from the
backlight.
2.2. Pixel Design
There are many variations in the pixel design layout. First, the designer must choose whether the a-Si:H TFTs will be
configured as back-channel-etched TFTs or channel-passivated TFTs (see Section 3.1). Next, the TFT layout and the
pixel electrode design must be consid

Figure 2.8 Two examples of pixel layouts: a) the additional capacitance type, Cadd, and b) the storage capacitance
type, Cst. These designs are not drawn to scale.

ered. Then, the configuration of the storage capacitance has to be decided. Two possible configurations are shown in
Figure 2.8. In the additional capacitance (C add) scheme (Figure 2.8 (a)), the pattern of the gate busline is formed to
overlap the transparent pixel electrode of ITO (Indium Tin Oxide), so that no extra buslines are needed. In the storage
capacitance design of C st (Figure 2.8 (b)), an independent electrode is provided for the storage capacitance. The C add
design is simple to fabricate, but it increases the gate busline capacitance. On the other hand, the C st design has an
increased number of cross-over points between the C st busline and the data busline; possible shorts between these
buslines can lower the production yield of TFT panels, and the panel processing becomes more complex.

The design rule determines the minimum pattern size and the gap or spacing between various patterns. For example, it
sets the minimum size for the spacing between pixel electrodes and data buslines, and also governs the channel length
of a-Si:H TFTs. A design rule or minimum design size of 10 um is generally adopted for panels with a diagonal size of
around 10" and video-graphic-array (VGA) resolution (640×480). For higher resolution panels, sub-10 um technology
is used for the design and processing.

As described before (Figure 2.4), the utility factor of the backlight illumination is limited by the aperture ratio of the
pixels. The aperture is defined as the area ratio of the transparent electrodes to the pixels. Since light can only be
modulated at the transparent pixel electrodes, light passing through the gaps between the pixel electrodes and the
metallization patterns degrades the display quality and should be eliminated. Therefore, the gap between metallizations
is covered by an opaque material; this is the so-called "black matrix" design. The black matrix is usually formed on the
color-filter substrate. Figure 2.9 shows an example of a black matrix layout and the color pixel (3 dots) layout. Since
the two glass substrates, i.e., the color-filter and TFT substrates, are assembled and aligned after processing each
substrate separately, the alignment cannot be as precise as can be achieved on a single substrate. Therefore, the black
matrix must have a large margin, resulting in a reduced aperture ratio. This large margin can be decreased by
fabricating the black matrix on the TFT substrate. Precise alignment can be achieved in this scheme, allowing the
aperture and brightness to be increased. However, since the materials for the black matrix are metals

Figure 2.9 Each pixel consists of three (R, G and B) dots. The black matrix is shown by the shaded area which also
cover the TFTs. In order to utilize effectively the black matrix area, the storage capacitance electrode (denser shape) is
extended vertically to form an H-shaped capacitance.

like chromium, capacitive coupling between the black matrix and the electrodes can be a problem. Therefore, the black
matrix is usually fabricated on the color-filter substrate. When chromium is used for the black matrix, it reflects light
from outside, degrading the display quality of the panel. This effect can be lessened by the use of a chromium oxide.
Another role of the black matrix is to shield the a-Si:H TFTs from incident light. As described in a later section, a-
Si:H is very sensitive to visible radiation. The visible light from the backlight incident upon the TFTs generates a
photocurrent that increases the TFT off-current. Thus, the black matrix is designed to cover the whole TFT area.

The aperture ratio is also affected by the disclination of the liquid-crystal. The disclination line corresponds to the
boundary between regions of the liquid-crystal that have directors with opposite orientations. In a TFT/LCD, the
electric field is generally applied vertically or perpendicular to the glass substrate. However, the electrode
configuration on the TFT substrate gives rise to a local horizontal electric field. This local field, in turn, induces
reverse tilt domains or disclinations. This phenomenon is also related to the rubbing direction of the alignment layer.
Generally, the rubbing direction is at 45° to the busline direction and the disclination tends to occur at the corner of
each pixel electrode corresponding to the tail of the rubbing direction. The reverse tilt introduces transmittance
irregularities and decreases the contrast ratio in that region. Therefore, in order to avoid undesirable effects, this area
also has to be covered by the black matrix, even though reverse tilt only occurs occasionally.
A cross-section of a TFT/LCD pixel is shown in Figure 2.10. The TFT, storage capacitor, and busline metallizations
are fabricated on the TFT glass substrate, and the color filter, black matrix and common electrode are fabricated on the
color-filter substrate. The surfaces of these substrates are coated with polyimide resin films and buffed with a fabric to
align the orientation of the liquid-crystal molecules. Liquid-crystal is injected between the TFT substrate and the color-
filter substrate. The TFT has a bottom-gate structure, sometimes called an inverted-staggered electrode structure. The
a-Si:H TFT is very sensitive to light and the intense illumination from the backlight must be shielded by the gate
electrode. Therefore, it is important that the a-Si:H island lies within the area of the gate electrode as shown in Figure
2.9 and 2.10. Light incident from the top surface of the color-filter substrate is shielded by the black matrix. This black
matrix also covers the gaps between the buslines and the pixel electrodes.

Figure 2.11 shows an equivalent circuit representation of a pixel. The distinction between the source and the drain
electrodes of the TFTs is usually quite clear. In an n-channel TFT such as an a-Si:H TFT, the drain electrode is biased
with a higher potential than the source electrode which is usually grounded. In the TFT/LCD panel, the signal voltage
fed through the data busline has an alternating polarity as

Figure 2.10 Cross-section of a TFT/LCD pixel. The polyimide film is rubbed with a fabric to make the liquid-crystal
molecules align in the proper direction. The diameter of the spacers defines the cell gap of the liquid crystal, which is
chosen to be ~5 um.

shown in Figure 2.12. Therefore, during a positive cycle of the signal voltage, the TFT electrode on the data busline
corresponds to the drain electrode. However, during a negative cycle of the signal voltage, the situation is reversed,
and the data busline electrode is biased to a lower potential with respect to the pixel electrode and thus corresponds to
the source electrode in its usual notation. For convenience, however, we call the electrode connected to the data
busline the drain and the electrode connected to the pixel electrode the source. According to this notation, the parasitic
capacitance between the gate and the source is C gs, which corresponds to the overlapping capacitance between the
source and the gate electrodes. This parasitic capacitance induces a dc voltage offset, /_\V, on the alternating voltage
applied to the liquid-crystal (Section 2.3.2). The potential of the common electrode of ITO
Figure 2.11 Equivalent circuit of a pixel, showing the drain and source electrode notations. V com and Vst are the
voltages of the common electrode and the storage capacitance electrode, respectively. The node potential V n represents
the potential of the pixel electrode.

Figure 2.12 Voltage waveforms of TFT/LCD. The signal voltage is ac with positive and negative cycles. A voltage
jump occurs due to the feed-through voltage from the gate pulse which results in a dc voltage offset. The common
electrode voltage Vcom is adjusted to this voltage offset.

on the color-filter substrate (V com ) is shifted from the earth potential by /_\V to compensate for the dc potential
generated by the parasitic capacitance, C gs, V com is sometimes driven in ac mode in order to decrease the absolute
amplitude of the data signal. The electrode of the storage capacitance, referred to here as V st, is biased to the ground
potential. If the capacitance is formed between the gate busline and the transparent electrode (C add case), V st becomes
equal to the potential of the gate busline.

2.3. Design Analysis


This section discusses three major issues of the TFT/LCD design: (1) the charging behavior of the pixel capacitance by
the a-Si:H TFT, (2) the dc voltage offset generated by the parasitic capacitance, C gs, and (3) the delay and distortion
of the gate pulse voltage or the gate delay.
2.3.1. Charging of Pixel Capacitance

As shown in Figure 2.11, the signal voltage is applied to the liquid-crystal cell through a TFT which acts as a voltage
switch, and this voltage controls the intensity of the illumination from the backlight. Following the notation of Figure
2.11, we derive a formula for the charging behavior of a liquid-crystal cell. The stored charge Q n at a node or pixel
electrode is given by

(2.2)

where V n is the node potential, V g is the gate voltage, C lc is the capacitance of the liquid-crystal, and the charging
current or drain current of the a-Si:H TFT, I d , is given by

(2.3)

Since the TFT is operated in the linear region of its I d -V d characteristics to charge the liquid-crystal cell, the drain
current is given by the formula (gradual-channel approximation)

(2.4)

(2.5)

where u n is the electron mobility in the channel, C i is the gate insulator capacitance per unit area, V d is the drain
voltage, V t is the threshold voltage of the TFT, and W and L are the width and length of the TFT, respectively.

Under the assumption of constant parameters we obtain from Equations (2.2), (2.3), (2.4) and (2.5)

(2.6)

where C px=C lc +C st+C gs. As will be discussed in Sections 3.2.2 and 5.2, both C lc and C gs depend on voltage and
thus time. However, these variations have little effect on the charging behavior. Therefore, the solution to Equation
(2.6) provides us with a good description of the charging behavior. If we set an initial value as follows (Figure 2.12)

(2.7)

the node potential V n is given by the following equation for the case of charging from a negative cycle to a positive
cycle:

(2.8)
where is a time constant 2 and a and b are dimensionless constants. 3 These constants are given by

(2.9)

(2.10)

and

(2.11)

Since Equation (2.8) corresponds to the transition from a negative to a positive cycle, V d in this equation is positive
and V no is negative as shown in Figure 2.12. The initial value, V no is defined as the node potential at the leading edge
of a gate pulse. At this point, the node potential V n jumps by /_\V, which is given by

(2.12)

where V a is the pulse amplitude of the gate voltage. V no is the value of V n after this jump takes place. V no is
therefore nearly equal to V d in the preceding cycle, or -V d , and in this negative cycle, the node potential is equal to -V
d -/_\V. The time constant of the pixel is given by

(2.13)

where R lc is the resistance of the liquid-crystal cell and R off is the off-resistance of the a-Si:H TFT. If is large
enough, the node potential will not decay from its initially charged value. In this case, the initial node potential is

(2.14)

and the constants of Equations (2.10) and (2.11) are simplified to

(2.15)

For the transition from a positive to a negative cycle, the same formula as above can be used with V d <0, and V no>0.
In this case, V no is again approximately equal to -V d including the effect of the voltage step /_\V due to C gs, if of
Equation (2.13) is larger than the frame time.
The simulated charging behavior of the node potential is shown in Figure 2.13. In this example, the time constant of
the charging process (Equation (2.9)) is calculated to be =5 us and 1 us for positive and negative cycles,
respectively, corresponding to charging times (95%) of 13.3 us and 4.3 us, respectively. The difference between the
two cases is fairly large, as is apparent from Equation (2.9), and in this example the falltime is 3.1 times faster than the
risetime.

From Equation (2.8), when t << this formula reduces to

(2.16)

Figure 2.13 The charging characteristics of the node. The positive cycle corresponds to the transition from a negative
to a positive cycle (Vd =8V), and the negative cycle corresponds to the reverse case (Vd =-8v). The numerical values
used in the calculation are as follows: Cpx=1pF, u n =0.5 cm2 /Vs, Ci= 20 nF/cm2 , W/L=5, Vg =13 V, Vt=1 V, Vd =±
8V, Vno= 8V.

where the coefficient a is given by


(2.17)

Substituting the numerical values shown in the caption of Figure 2.13 into Equation (2.17) yields a=-6 and -1.2 for the
positive and negative cycles, respectively. Therefore, the charging speed at the onset of the positive cycle is fairly high
but it gets slower as it proceeds.

One of the most important parameters in designing a TFT is the ratio of the width to the length, W/L, of the transistor.
It is directly related to the time t r required to charge the pixel capacitance. If the ratio of the charged node potential to
the data voltage is defined as r c, we obtain from Equation (2.8)

(2.18)

where k t is given by

(2.19)

This must be shorter than the scan time, which is the frame time t f divided by the number of scan lines, m, and this
condition is given by

(2.20)

From Equations (2.9), (2.18), (2.19) and (2.20), we obtain

(2.21)

With a frame rate of 60 Hz, 480 scan lines, and r c=0.95, Equation (2.21) reduces to

(2.22)

Here, we assumed the same figures as described above. This value for W/L is almost ideal from the viewpoint of TFT
design, since W/L=1 provides the minimum TFT cell area. If we assume r c=0.99, however, Equation (2.21) reduces to
W/L>3.0.

In the derivation of Equation (2.22), we assumed the gate line delay associated with voltage pulses travelling along the
gate busline was negligible. However, the gate delay can represent a considerable part of the scan time as TFT/LCD
panels get larger and their resolution gets higher. When the gate delay is included, Equation (2.20) becomes

(2.23)

and numerical evaluation of this equation with r c=0.99 and t d =10 us gives
(2.24)

which is still a reasonable value.

Another basic criterion can be deduced from the risetime formula of Equation (2.20). The TFT must have sufficient
retention capability so as not to discharge the stored charge during the frame time. The retention ratio r is written as

(2.25)

If of Equation (2.13) is assumed to be much larger than t f , we obtain from Equations (2.18), (2.19), (2.20), and
(2.25):

(2.26)

If the on-resistance of the TFT R on is defined as

(2.27)

the ratio of the on- and off-resistance of the TFT is obtained from Equations (2.9), (2.13), (2.26) and (2.27):

(2.28)

For 480 scan lines, a retention ratio of 99% (r=0.99), r c=0.99, and R lc R off , the on/off ratio of the TFT is given by

(2.29)

where we assumed the same voltage figures as in Figure 2.13. This criterion sets a minimum level of the on/off current
ratio of the TFTs. Although the transfer characteristics of the a-Si:H TFT seem to satisfy this condition as will be seen
in Section 3.2, the numerical check will be of value. By substituting u n =0.5 cm2 /Vs, C i=20 nF/cm2 , W/L=5, Vg = 13
V, V t=1 V, and V d =8 V into Equation (2.27), we obtain an on-resistance of 5×10 6 n. With this, R on, R off is
calculated to be 2×10 12 n from Equation (2.29). These resistance values correspond to on- and off-current of 2 uA and
5 pA for V d =10 V, respectively, and this current range can well be covered by a-Si:H TFTs. From an assumption of R
lc R off , the resistivity of liquid-crystal is calculated to be 5×10 11 n cm for a cell gap of 5 um and an aperture of
250×50 um 2 in a pixel area of 300 ×100 um 2 . Since the bulk resistivity of liquid-crystal is 1012–1014 n cm, there is a
wide margin. The time constant of the pixel is calculated from Equation (2.13) to be one second, which is also
consistent with the assumption made in deriving Equation (2.26) from Equation (2.25).

2.3.2. Voltage Offset

Liquid-crystal displays do not work properly under dc bias, so they are always operated in ac mode as shown by V d in
Figure 2.12. Due to this ac operation and the parasitic capacitance formed between the gate and the source electrode of
the TFT, however, a dc voltage offset, /_\V, appears in the node potential, V n 4 . This offset is given by

(2.30)

where V a is the amplitude of the voltage pulse applied to the gate busline to turn on the TFTs connected to it. When
the gate voltage appears on the line, C gs is charged at the leading edge and discharged at the trailing edge of the
voltage pulse. Under normal conditions, the charging and discharging cancel each other out, and no effect is observed.
However, in TFT/LCD panels, the ac signal is applied to the liquid-crystal through a TFT, and only the discharging
effect remains on the pixel electrode as shown in Figure 2.12.

The resultant dc voltage offset causes undesirable effects on the performance of the liquid-crystal display such as
flicker 5 , image sticking 6 , 7 and permanent brightness nonuniformities on the panel. Flicker noise appears as low-
frequency brightness variation. The voltage offset causes asymmetry in the ac node potential from frame to frame. The
liquid-crystal cell reacts to this asymmetrical signal voltage, resulting in fluctuation of its transmittance. When the
frame frequency is 60 Hz, this fluctuation occurs at a frequency of 30 Hz, and is recognizable.

The storage capacitance is included in the design of the pixel electrode to reduce this level shift. Without C st,
Equation (2.30) reduces to

(2.31)

For V a=22 V (assumed base level of gate voltage is -9V), C gs=0.1 pF, and C lc =0.3 pF, /_\V is as high as 5.5 V.
Since the threshold voltage, V th , of a TN liquid-crystal is 2–3 V, this value of /_\V is about twice that of Vth . With C
st, this voltage shift takes a lower value. If the same figures are substituted as described above with C st=0.6 pF, /_\V
becomes

(2.32)

This represents a considerable improvement, but the dc bias is still rather high. This is compensated by adjusting the
voltage of the common electrode on the color filter substrate, V com , so as to cancel out the voltage shift. However, this
canceling is not perfect since both C lc and C gs of Equation (2.30) are voltage dependent.

The parasitic capacitance depends on V g as will be discussed in Section 3.2.2. However, since the transient of the gate
pulse (~ns), is much faster than the response time of a-Si:H TFTs (~us), C gs can be assumed to be constant and be
approximated by the value corresponding to the peak gate voltage. Equation (2.30) then reduces to

(2.33)

As will be described in Section 5.2, the liquid-crystal capacitance varies as a function of applied voltage. This
variation is due to the anisotropic dielectric constant of the liquid-crystal. In a TN liquid-crystal, the dielectric constant
is low ( = - ) when there is no applied voltage and approaches // as the voltage applied to the cell is increased.

The difference in /_\V between two pixels, dV, is given by 6


(2.34)

where V 1 and V 2 are the applied rms voltages to these pixels. The maximum value of Equation (2.34), is obtained by
substituting C lc (V 1 ) and C lc (V 2 ) with C - . and C // , which correspond to = - and = // , respectively. When C-
=0.3 pF and C // =0.6 pF, /_\V is 2.2 V and 1.7 V. Therefore V com is set to -1.95 V in this case, and dV is 500 mV. The
offset voltage dV is given by dV=±250 mV where

(2.35)

If there is no parallel (storage) capacitance, dV=±1.2V when V com = -4.3 V. Therefore, in this case the effect of the
storage capacitance is to decrease dV by a factor of 4.8.

Although the display quality is improved considerably by the addition of a storage capacitance (C st or C add), there
still remains a residual dc voltage of 250 mV. Increasing the storage capacitance is the simplest way to lower /_\V, but
this reduces the aperture ratio since the area taken up by the storage capacitance will increase. Another approach is to
decrease C gs by decreasing the overlap area between the source and the gate electrode of the TFT. This can be
achieved with self-aligning technology 8 , 9 , in which the gate electrode is used as a mask to define the amorphous
silicon island and the edge of the passivation layer. The glass substrate is illuminated from below. A typical example
of self-aligned TFT processing will be shown in Figure 3.26. If the a-Si:H TFT is fabricated with a perfect self-
alignment process, i.e., with little overlap between the gate and source electrodes, it does not work well. A small
overlap area between the source and gate electrodes impedes the injection of carriers, resulting in a high-resistance
region that impairs the TFT operation; an overlap width of about 1 um is necessary to achieve sufficient injection.

The effect of reducing C gs can be estimated as follows: if the parasitic capacitance is decreased from C gs to C gs/m
where m is a constant larger than 1, /_\V in Equation (2.31) is given by

(2.36)

which reduces to

(2.37)

Although this is nominally equivalent to substituting C st=(m-1) C lc in Equation (2.30), the effect is lessened due to
the voltage dependence of C lc . Calculation of dV shows that the effect is equivalent to adding a storage capacitor of
only 0.9 C - and 1.2 C - for m=3 and m=4, respectively, where we reasonably assumed that C - =3 C gs and C // =2C - .
Therefore, reduction of C gs is not so effective to lower dV as expected.

2.3.3. Gate Delay

TFT/LCD panels like the one shown in Figure 2.6 are operated on a line-at-a-time basis. Each gate busline is selected
sequentially and a voltage pulse is applied to the selected gate busline. This pulse propagates down the busline which,
we assume, is represented by the CR
Figure 2.14 Distributed-constant representation of the gate busline. The voltage pulse is delayed and distorted by the
series resistance and parallel capacitance of the line as it propagates along the gate busline. R and C correspond to the
resistance and capacitance per unit length, respectively.

distributed-constant circuit shown in Figure 2.14, where C and R are the capacitance and resistance per unit length.
The rectangular gate pulse at the input is delayed and distorted as it travels along the gate busline. The gate delay is
defined as the time between the onset of the pulse at the starting point of the busline and the time when the pulse
height at the end of the busline reaches 90% of the pulse amplitude Vg .

The transmission of pulses through the distributed-constant RC line is governed by the following equations:

(2.38)

(2.39)

where the voltage v and the current i are both functions of x and t. This is the so-called Thomson cable equation,
which has been analyzed for the cases of an infinite line and finite-length lines terminated with open circuits and short
circuits. The TFT/LCD gate busline corresponds to the case of a finite-length RC line terminated with an open circuit.
In this case, Equations (2.38) and (2.39) are reduced to

(2.40)

(2.41)

The Laplace transform is a convenient tool for analyzing this problem. The unilateral Laplace transform of v and i is
given by

(2.42)
(2.43)

From the initial condition of v=0, i=0 at t<0 we obtain

(2.44)

(2.45)

(2.46)

(2.47)

The solution of Equation (2.46) is given by

(2.48)

Substituting Equation (2.48) into Equation (2.45), we obtain

(2.49)

Then,

(2.50)

Taking into account the fact that I and V satisfy the same type of partial differential equation, we obtain C(x)=0, and
Equation (2.50) reduces to

(2.51)

where A(s) and B(s) are independent of x and are determined from the boundary conditions.

Equations (2.48) and (2.51) are the basic equations of an RC line. For the case of infinite RC line, the solution is given
by an error function. However, this will not be discussed in detail here. When the RC line has finite length L g and
open-circuit termination, the boundary conditions are
(2.52)

(2.53)

where F(s) is the Laplace transform of a given voltage function f(t) at the input (x=0). From Equations (2.48), (2.51),
(2.52) and (2.53),

(2.54)

(2.55)

From Equations (2.48), (2.51), (2.54) and (2.55),

(2.56)

(2.57)

For indicial (step) response, F(s)=V g /s, and Equation (2.56) reduces to

(2.58)

The singular points of Equation (2.58) are s=0 and the roots of cosh =0. The solutions are given by

(2.59)

The inverse Laplace transform yields the solution of v(x, t) and the corresponding i(x, t) as follows:

(2.60)

From Equations (2.57) and (2.60),


(2.61)

The gate delay is estimated by putting x=L g in Equation (2.60), and is given by

(2.62)

When t/RCL g 2 >1, Equation (2.62) is approximated as

(2.63)

The gate delay t d of TFT/LCD panels can be estimated from (2.63):

(2.64)

where t d is defined as the time at which the voltage reaches 90% of the pulse voltage. Therefore, t d can be closely
approximated by a simple expression of RCL g 2 . Another good approximation for t d is 2 RCL g 2 , which corresponds
to the time for 99% charging. A more practical expression for t d (90%) is R p C p n p 2 where R p and C p are the
resistance and capacitance per pixel (dot), respectively, and np is the number of pixels (dots).

In Equation (2.64), R is the resistance per unit length of the gate metallization. In small- and medium-sized panels, Cr
is the most widely used metal. With Cr's specific resistivity of 55 un cm and SiN gate insulators with a dielectric
constant of 6.9, t d becomes as high as 20 us. This is adequate for low- and medium-resolution panels in which the
gate selection time is (1/60)×(1/480)=34.7 us or longer. However, in high-resolution panels with over a thousand gate
lines, the gate address time is 16.7 us or less. In this case the gate metallization must be made with lower resistivity
metal. Al gate metallization, as will be described in Section 3.4.1, is suitable for this purpose. The resistivity of Al is
3–6 un cm, and from Equation (2.64) the value of t d is 1–2 us for the case of 10-inch (25-cm) TFT/LCD panels. Since
the charging time of a unit cell of liquid-crystal is approximately 13 us, as discussed in Section 2.3.1, t d of 1–2 us is
more than reasonable.

The diagonal size is limited by the resistivity of the metals used for the gate line metallization. The lower the
resistivity, the larger the panels that can be fabricated and the higher their resolution. The calculated results are shown
in Figure 2.15 (a) and (b). Larger panels can be designed with C st, since the dotwise capacitance is larger in C add case
than in C st case due to a large overlap area of C add between the gate
Figure 2.15 Relation between the gate metal resistivity and the diagonal size of the display panel: (a) the panel size of
VGA displays where the number of pixels is 640×480; (b) the size of higher-resolution panels (1920×1120) 10 . when
aluminum is used for the gate busline metallization, panels larger than 30" can be fabricated, even those of the high-
resolution type.

busline and the pixel electrode. The maximum diagonal size can reach 30 inches or more for Al metallization.
According to these results, a-Si:H TFTs are applicable to a wide range of TFT/LCD panels. Figure 2.15 shows the
results calculated with a rather conservative design rule, i.e., assuming a gate busline width of 10 um. Different designs
including wider gate buslines and larger W/L of a-Si:H TFTs will allow panels larger than one metre (40") to be
constructed.

2.3.4. Colorimetric Design

The colorimetric system of the TFT/LCD is designed according to the chromaticity diagram recommended by the 1931
CIE (Commission Internationale de l'Eclairage) conference. Any color can be matched by additively mixing the three
primary colors according to the spectral tristimulus values or color matching functions obtained for all visible
wavelengths. The average of the tristimulus values of many observers are represented by ( ), ( ), and ( ). The
CIE-RGB standard colorimetric system corresponds to the observation of color with a viewing angle of two degrees
and is called the XYZ colorimetric system. Three monochromatic sources of wavelengths 700.0, 546.1, and 435.8 nm
were chosen as reference stimuli, and the values of these stimuli were determined to represent a specific radiation of
one watt per nanometer. The tristimulus values of the CIE-RGB colorimetric system are shown in Figure 2.16.

The XYZ colorimetric system is a system in which ( ), ( ), and ( ) are three color matching functions that are
transformed from ( ), ( ), and ( ) as follows:

(2.65)

The tristimulus values of the XYZ colorimetric system are shown in Figure 2.17. This system is suitable for practical
applications since it does not involve negative values. The tristimulus values of the light-source color are given by

(2.66)

Figure 2.16 Color matching function or spectral tristimulus values of the CIE-RGB colorimetric system. A stimulus
value of less than zero is added to the radiation to be matched.

where k is a coefficient, P( ) is the spectral energy distribution, and integration is performed over the visible range.
When P( ) is given as the special radiance, Y represents the luminance. The coefficient k is usually normalized as

(2.67)

The chromaticity coordinates x, y, and z are defined as the ratio of the tristimulus values to the total stimulus
S(=X+Y+Z) as follows:

Figure 2.17 Color matching functions of the CIE-XYZ colorimetric system.

(2.68)

Because x, y and z are connected by the relation x+y+z=1, two of the three variables are independent, and the (x, y)
coordinate system is usually adopted to represent the chromaticity diagram as shown in Figure 2.18, which shows the
projection of the x+y+z=1 plane on the x-y plane. The locus of pure spectral radiation is shown on the horseshoe curve
with numerical values showing the wavelength in um. The line connecting the edges of the spectral locus is called the
purple

Figure 2.18 CIE chromaticity diagram. Monochromatic radiation is shown on the spectral locus where the wavelength
is shown in microns. The Planckian locus corresponds to the black-body radiation of a specific color temperature.

boundary. The colors on this line correspond to mixtures of red and purple. Visible colors are represented by points in
the region bounded by the spectral locus and the purple boundary.

Additive mixtures of color stimuli can be displayed on the chromaticity diagram by connecting the chromaticity points
P 1 and P 2 of the two color stimuli to be mixed. If the chromaticity coordinates and the total stimuli of the
chromaticity points of P 1 and P 2 are written as (x 1 , y 1 , S 1 ) and (x 2 , y 2 , S 2 ), respectively, the chromaticity
coordinate (x, y) and the total stimulus S of the mixed color chromaticity point P are given by

(2.69)

P is on the line connecting P 1 and P 2 and the distances are given by the relation

(2.70)
Figure 2.18 also shows the Planckian locus, which corresponds to the chromaticity of black-body radiation. The
spectral density distribution of the black-body radiation P( ) at the thermal equilibrium of temperature T, is given by

(2.71)

where C 2 is the secondary radiation constant (C 2 =1.4388×10 -2 mK) and C 1 is an arbitrary constant. The CIE defines
four standard illuminants denoted by A, B, C, and D, which are shown in Figure 2.18. Standard illuminant C
corresponds to average daylight with a color temperature of 6774K. When the R, G and B stimuli of an NTSC
(National Television Standard Committee) signal are all the same, the mixture of these colors yields the standard
illuminant C of which the chromaticity coordinates are x=0.310, y=0.316. The three primary colors of NTSC are
shown in Figure 2.20.

The three color stimuli of X,Y, Z or the TFT/LCD panels are given by

(2.72)

where S( ) is the spectral distribution of the backlight illumination, T( ) is the transmittance of the liquid-crystal cell
and the color filter, and k is a coefficient. The coefficient k is normalized by

(2.73)

The spectral distribution of the backlight is shown in Figure 2.5. This sort of three-wavelength illumination is most
widely used. The transmittance of the color filters as a function of wavelength is shown in Figure 2.19. The most
common material used to fabricate the color filter is gelatin. A gelatin film about 1–2 um thick is coated on the glass
substrate and then photolithographically patterned. The material is then dyed to a specific color and the surface is
coated with a passivation layer. This series of process steps is repeated three times for R, G and B colors to complete
the color filter fabrication. Another way of fabricating the color filter is based on patterning of a pigment-dispersed
resin. The feature of this method is the stability against heat (260°C)
Figure 2.19 Transmission spectra of color filters made of dyed gelatin.
Figure 2.20 Comparison of the color coordinates of a TFT/LCD with those of NTSC signals and CRT displays. The
coordinates of NTSC (National Television Standard Committee) signals are R=(0.67, 033), G=(0.21, 0.71), and
B=(0.14, 0.08).

and against intense light exposure. The combined characteristics of the backlight illumination and the color filter
transmittance determine the spectral properties of the TFT/LCD, i.e., the color reproduction characteristics. The color
coordinates of a typical TFT/LCD panel are shown on the CIE chromaticity diagram in Figure 2.20. The range of
colors covered by the TFT panel is comparable to that of the CRT.

In TFT/LCDs, each pixel is composed of R, G and B color dots as shown in Figure 2.21. The triad and quad
arrangements shown in (b) and (c) are often used in TVs. There is also a variant of the quad

Figure 2.21 Color-dot arrangements of TFT/LCDs. The striped pattern is usually used in computer displays, while the
triad or quad types are used in TVs.
arrangement in which one of the two green dots is replaced by a white dot. In computer terminal displays, however,
the striped three-dot arrangement (a) is generally used. The digital signal is fed to each dot, and it is possible to display
(2 n ) 3 colors for n bits per color channel. For n=2, the panel display 64 colors, and 4,096 colors for n=4. The 24-bit
color graphics represent 16,777,216 colors (=(28 ) 3 ).

2.4. TFT Panel Fabrication


The fabrication of a-Si:H TFTs is based on thin-film technology, i.e., thin-film deposition and thin-film pattern
etching. This technology is a natural extension of crystalline silicon LSI (large-scale integration) technology, although
there are some differences between the two. In LSIs, the substrate is a silicon wafer on which a variety of processing
steps are performed, such as surface finishing, rinsing, thin-film coating, thin-film deposition, thin-film growth,
oxidation, etching, lithography, impurity diffusion, heat treatment, and ion implantation. Silicon is quite stable under
extreme conditions which means that processes like oxidation, diffusion, and heat treatment can be carried out at
temperatures higher than 1000°C (T m =1420°C).

However, the TFTs have to be processed on transparent glass substrates, so that the process temperature is limited to
below, say, 400°C. Amorphous silicon-based technology has an advantage over other technologies in that the
processing can all be carried out below 350°C. The process steps include thin-film deposition such as plasma CVD
(chemical vapor deposition) and metal sputtering, thin-film coating, lithography, oxidation, etching, and rinsing.

The substrate size in TFT fabrication also differs from that of LSIs. LSI wafers are only a few inches in diameter, 6"-
8" being the most popular sizes. TFTs, on the other hand, have much larger substrates that are rectangular in shape,
having diagonal sizes ranging from a few inches to a few feet. Substrates with diagonals of over 20" are now used in
the production lines of TFT/LCDs.

Such large substrates require large optical lithography systems. Both one-to-one projection and step-and-repeat
aligners are used in the photolithographic systems for TFT fabrication. The step-and-repeat aligner shown in Figure
2.22 features accurate alignment and fine patterning,
Figure 2.22 Configuration of a step-and-repeat aligner.* This system covers the whole surface of the substrate by
changing reticles.

*Nikon Step and Repeat System for LCDs

while the one-to-one projection aligner features high throughput. The typical specifications of current steppers are
shown in Table 2.1. The typical design rule for 10" class TFT panels is 10 um, which corresponds to the minimum
width of the gate or data buslines. To achieve fine patterning, a positive photoresist is generally used, where the
exposed part of the photoresist is etched away.

Since any dust in the clean room has the potential of making a panel defective, great care must be taken to get rid of
particles. The rinse process is important to produce panels without any point (line) defects. Almost one third of the
total processing is devoted to rinsing the panel. The TFT panel process is carried out on a glass substrate. Glass has
suitable attributes for flat-panel displays as well as conventional CRTs, i.e., transparency, rigidity, and thermal
stability. In addition to this, the TFT-LCD substrates must also be flat, free from surface and internal defects and
scratches, resistant to temperature cycles and a variety of chemical etchants, and must have low alkali content.

The most widely used and accepted substrate glass is Corning 7059 glass which is made by the fusion down draw
machine. 12 The composition of 7059 glass is shown in Table 2.2. A stream of homogeneous molten glass is delivered
into a tapered trough at the top of a refractory form called a fusion pipe. When the trough overflows, the glass flows
down in sheets on both sides of the fusion pipe. Two glass sheets are manufactured at the same time, and the surfaces
of these sheets do not contact any other surfaces during the forming process. In this configuration, glass with a width
of approximately 40 inches can be manufactured.

The thermal stability of this glass, which is essentially a function of glass viscosity, is very good. The strain point of
7059 glass, defined as the temperature at which the viscosity is approximately 1014.5 Poiseulles,

is ~593°C, while its softening point (viscosity=10 7.6 Poiseulles) is ~844°C. The maximum temperature used in the
fabrication process is the strain point minus 25°C, which leaves a considerable margin for the amorphous silicon,
which is processed below 350°C. In this temperature range, the thermal shrinkage and warp are small since they are
related to glass viscosity. The chemical durability of 7059 glass can be estimated from the weight loss under severe
etching conditions: when it is dipped in a 10% HF solution for 20 minutes at room temperature, it loses approximately
10 mg/cm 2 . This is quite acceptable for the amorphous silicon based panel fabrication process.

The TFT process has to satisfy various requirements since it is a thin-film, low-temperature, large-area, and fine-
pattern process. The cell gap (the thickness of the liquid-crystal cells) is around 5 um. Thick-film processing is not
compatible with such a small gap and the TFT process has to be a thin-film process. The requirements for low
temperature result from the substrate being a transparent glass, and fine patterns are required for high resolution. An
outline of the process flow is shown in Figure 2.23, which corresponds to the TFT shown in Figure 3.4 (a). The process
is essentially the same as TFT fabrication. However, in addition to this, the following parts must be fabricated: gate
and data buslines, the storage capacitor, and transparent pixel electrodes.

The transparent pixel electrodes are made by sputtering a target of indium-tin-oxide (ITO) and then patterning it. In a
vacuum chamber, the target is sputtered in an atmosphere of argon under a pressure of mTorr. The ITO film is then
patterned in an aqueous solution of nitric acid and chloric acid. The ITO film is compatible with thin-film processing
and its transmittance is higher than 90% for the light in the visible spectrum and its resistivity is about 10-4 n cm.
Figure 2.24 shows the ranges of pressure and temperature used to deposit various thin films. Table 2.3 and Table 2.4
show the conditions under which the films are deposited and etched.

In Figure 2.23, four mask steps are shown:

1. The gate electrode of the TFT and the gate busline,

2. the a-Si:H island,

3. the pixel electrode, and

4. the source and drain electrodes and the data busline.

In addition to this, the final passivation layer is deposited and pattern etched. Therefore, five is the standard number of
mask steps in the
Figure 2.23 Outline of the process steps involved in fabricating TFT panels. Each step corresponds to a
photolithographic mask step. Since the final passivation process is omitted, five mask steps are necessary.
Figure 2.24 Pressure and temperature ranges for the deposition of thin films during TFT panel fabrication.

panel process corresponding to the back-channel etched TFT. The minimum pattern size in these steps is determined
from such factors as the panel size, the resolution, the aperture, and the productivity. The other design parameters to
consider related to the process are the line width, the overlapping width, and separation between different patterns in
different masks. These parameters are summed up as a design rule in a particular process line and are observed
throughout the whole panel process.

Figure 2.25 shows the process flow of the TFT panel which corresponds to the TFT shown in Figure 3.4 (b). Due to
the introduction of the channel-passivation layer on top of the a-Si:H layer, one more mask is necessary to fabricate
the panel compared with the process described
above for the BCE TFT. For the formation of the ohmic contact layer, micro-crystalline n+ silicon is sometimes used
instead of heavily doped n+ hydrogenated amorphous silicon.

References
1. Inoue, F., Ando, K., Kabuto, N., Kamiya, M., Nakatani, N., Tsukada, T., et al. (1988) A 5-in.-diagonal liquid-
crystal color TV. In Digest of Technical Papers of the Society for Information Display International Symposium
(Anaheim, 1988), pp. 318–321, California:SID.

2. Kaneko, Y., Sasano, A., and Tsukada, T. (1989). Analysis and design of a-Si TFT/LCD panels with a pixel model.
IEEE Transaction Electron Devices , ED-36, 2953–2958.

3. Tsukada, T. (1993). State-of-the-art of a-Si TFT/LCD. Transaction of the Institute of Electronics, Information and
Communication Engineers (Japan), J76-C-II, pp. 177–183.
Tsukada, T. (1993). Development of aluminum gate thin-film transistors based on aluminum oxide insulators. In
Amorphous Insulating Thin Films. Material Research Society Symposium Proceedings (Boston, 1992), edited by
J.Kanicki et al., pp. 371–382. Pittsburgh: Material Research Society.

4. Morin, F. (1983). Electrooptical performance of a TFT-addressed TNLC panel. Proc. 3rd International Display
Research Conference (Kobe, 1983). pp. 412–414, California:SID.

5. Kaneko, Y., Tanaka, Y., Kabuto, N., and Tsukada, T. (1989). A new address scheme to improve the display quality
of a-Si TFT/LCD panel. IEEE Transaction Electron Devices , ED-36, 2949–2952.

6. Nanno, Y., Mino, Y., Takeda, E., and Nagata, S. (1990). Characterization of sticking effects of TFT-LCD. In Digest
of Technical Papers of the Society for Information Display International Symposium (Las Vegas, 1990), pp. 404–407.
California: SID.

7. Kanemori, Y., Katayama, M., Nakazawa, K., Kato, H., Yano, K., Fukuoka, Y., et al. (1990). 10.4-in-diagonal color
TFT-LCDs without residual images. In Digest of Technical Papers of the Society for Information Display
International Symposium (Las Vegas, 1990), pp. 408–411. California: SID.
Figure 2.25 Outline of the process steps to fabricate a CHP TFT panel. Six mask steps are necessary to complete the
process.
8. Nasu, Y., Kawai, S., Kisumi, S., Oki, K., and Hori, K. (1986). Color LCD for character and TV display addressed
by self-aligned a-Si:H TFT. In Digest of Technical Papers of the Society for Information Display International
Symposium (San Diego, 1986), pp. 289–292. California:SID.

9. Asama, K., Kodama, T., Kawai, S., Nasu, Y., and Yanagisawa, S. (1983). A self-alignment processed a-Si TFT
matrix circuit for LCD panels. In Digest of Technical Papers of the Society for Information Display International
Symposium (Philadelphia, 1983), pp. 144–145. California: SID.

10. Yamamoto, H., Matsumaru, H., Tsutsui, K., Konishi, N., Nakatani, M., Shirahashi, K., Sasano, A., and Tsukada, T.
(1990). A new a-Si TFT with Al2 O3 /SiN double-layered gate insulator for 10.4-inch diagonal multicolor display. In
Technical Digest of the International Electron Devices Meeting (San Francisco, 1990), pp. 851–854. New York: IEEE.

11. Ikeda, N., Moriyama, H., Uchida, H., Nishida, S., Mitsuhashi, K., Matsuo, O., et al. (1992). High-resolution 12.9-
in. multicolor TFT-LCD for EWS. In Digest of Technical Papers of the Society for Information Display International
Symposium (Boston, 1992). pp. 59–62, California: SID.

12. Dumbaugh, W.H. and Bocko, P.L. (1990). Substrate glasses for flat-panel displays. In Digest of Technical Papers
of the Society for Information Display International Symposium (Las Vegas, 1990), pp. 70–72. California: SID.
CHAPTER 3
Thin-Film Transistors
Thin-film transistors (TFTs) were proposed by Weimer 1 , 2 in 1961. The semiconductor layer and gate insulator of the
original TFTs were made of cadmium sulfide and silicon monoxide, respectively. These thin-film layers were
deposited in vacuum by evaporation. In spite of the complexity of fabricating TFTs in a vacuum chamber, they
obtained both depletion-type and enhancement-type TFTs with good saturation-current characteristics. At first, they
applied these TFTs to thin-film logic circuitry and built circuits such as flip-flops, AND gates, and NOR gates for
computer applications. Soon after this, the application of TFTs to liquid-crystal displays was also proposed and
various materials have been studied for this purpose. Among them, CdSe TFTs 3 , 4 have been intensively studied, and
TFT-addressed liquid-crystal panels of up to 6"×6" in size have been fabricated. Such panels are typically made with
Al or Au as metallization, CdSe as the semiconductor, and SiO2 or Al2 O3 as the insulator. Their applicability to
alphanumeric and video displays has been demonstrated on a 180×180 pixel panel.

Hydrogenated-amorphous-silicon thin-film transistors (a-Si:H TFT) were first reported in 1979 5 . The transfer
characteristics of an a-Si:H TFT with a gate insulator of SiN are shown in Figure 3.1. 6 The transfer characteristics
correspond to the behavior of the drain current, I d , as a function of the gate voltage, Vg . As can be seen, the drain
current has a wide dynamic range with an on-current in excess of 10- 6 A and an off-current below 10- 11 A. The
output characteristics of this TFT (i.e., the relation between the drain current, I d , and the drain voltage, V d are shown
in Figure 3.2. Saturation of the drain current is clearly demonstrated. The threshold voltage, V t, of the device was ~5 V
and the transconductance, g m , was measured as gm =0.3 un-1 at V d =10 V and V g =20 V. The field effect mobility
was estimated to be 0.4 cm2 /Vs from the transconductance. As a matter of fact, the transfer characteristics of a-Si:H
TFTs was implied back in 1972 7 by Spear and LeComber, who used field effect techniques to determine the
distribution function
Figure 3.1 Transfer characteristics of an a-Si:H TFT 6 . The on/off ratio of the drain current is about 106 . The sharp
subthreshold slope is caused by a strong accumulation layer formed at the interface between the hydrogenated
amorphous silicon and the silicon nitride.

of localized states in a-Si:H. Figure 3 of their paper clearly demonstrates the TFT transfer characteristics, although the
potential was in the order of 1,000 V. The gate insulator was a glass substrate with a thickness of 250 um.

Although liquid-crystal displays were the original application target of a-Si:H TFTs, it was quite a while before the
development of an a-Si:H TFT liquid-crystal display was first reported 8 . This was a 96×96 mm2 display containing
240×240 dots. Since then, a-Si:H TFTs have been widely used to drive active-matrix LCDs.
Figure 3.2 Output characteristics of an a-Si:H TFT 6 . The TFT channel is 500 um wide and 40 um long, and the a-
Si:H and SiN layers are both 0.5 um think. Therefore, the operating voltages can be made lower than those shown
here, as was later reported.

3.1. Hydrogenated Amorphous Silicon Thin-Film Transistors


The a-Si:H TFT is now widely recognized to be the most important and successful active device for use in active
matrix liquid-crystal displays. Three-terminal devices (like TFTs) are more flexible in their operation and have fewer
limitations than two-terminal devices (i.e. diodes). However, this is not the only reason for using a-Si:H TFTs as the
active devices to drive liquid-crystal cells. Hydrogenated amorphous silicon is a material with well-balanced features
for electronic applications.

Some of the advantages of a-Si:H are as follows:

1. It can be deposited over a large area: a-Si:H is deposited by plasma chemical vapor deposition (p-CVD) in which
silane gas (SiH 4 ) is decomposed in a plasma excited by rf power. A typical rf frequency is 13.56 MHz, although much
lower frequencies (e.g. 100 Hz) can also be used. The deposition system is shown schematically in Figure 3.3. The
diode type reactor contains two electrodes, on one of which the substrates are placed. After the chamber is evacuated,
silane gas, hydrogen diluting gas and a doping gas such as phosphine or diborane are introduced into the chamber. The
rf power is then
Figure 3.3 A plasma CVD system for depositing thin films of hydrogenated amorphous silicon and silicon nitride. The
silane gas is dissociated in an rf chamber to deposit the a-Si:H film, and the ammonia gas and nitrogen gas are added
to deposit SiN. To dope a-Si:H, phosphine (PH3 ) or diborane (B2 H5 ) gas are introduced into the chamber.

applied at a gas pressure of 0.1–1 Torr. The plasma is confined to the space between the two parallel electrodes. The rf
power is distributed uniformly over the surface of the electrodes, whose diameter exceeds one metre. Using this
technique, films can be deposited with thickness variations of only a few percent over the entire surface area. This
p-CVD process is also called plasma enhanced CVD (PECVD).

2. It can be deposited at low temperature: Low temperature deposition is essential in the fabrication of active matrix
liquid-crystal displays because glass is used as the transparent substrate material. In the p-CVD system, the silane
molecules are dissociated by an rf plasma. Consequently, the substrate temperature can be kept low in contrast to
thermal CVD systems in which the silane is thermally dissociated. The deposition temperature must be low because
good-quality a-Si:H films can only be formed if hydrogen is incorporated into the amorphous network; hydrogen will
not be retained in the film if the substrate temperature is higher than 450°C.

3. a-Si:H is an amorphous material, so it can easily form hetero-interfaces while maintaining good interface properties.
It is possible to deposit amorphous silicon on various substrates such as insulators (including glass, oxides, and
nitrides), metals and semiconductors. Especially important is the interface between silicon nitride and a-Si:H. The
interface properties between the a-Si:H and the SiN gate insulator play a critical role in TFT characteristics. Metals
like Cr, Al, Ta, and Mo are used as the metallization material and ITO (Indium Tin-Oxide) is also used as the
conductive film. The a-Si:H forms good interfaces with all these materials, making the processing of TFTs very
flexible and allowing a variety of processing schemes to be devised and used in the process line.

4. Amorphous silicon is a hard scratch-resistant material (Vickers hardness HV=1,500–2,000 kg/mm 2 ) and can be
finely patterned with photolithographic technology. A minimum dimension for 10" diagonal panels of about 10 um can
be achieved without difficulty—the limit of the fine pattern lithographic process of a-Si:H will be well below the
submicron level.

5. Hydrogenated amorphous silicon has high electrical resistivity in its undoped state and is highly conductive when
doped. The high resistivity of a-Si:H matches the high resistivity of the liquid-crystal,
Figure 3.4 Cross-sections of inverted staggered-electrode a-Si:H TFTs with (a) back-channel-etched and (b) channel-
passivated configurations. The black areas on top of the a-Si:H layers are ohmic contact layers of n+ a-Si:H.

i.e., the off-state resistance of a-Si:H TFT is comparable to the resistance of the liquid-crystal cell.

6. Non-toxicity is another favorable feature of a-Si:H.

The inverted staggered-electrode structure is most widely used in TFT/LCD fabrication. Figure 3.4 (a) shows a cross-
section of the back-channel-etched (BCE) TFT, and Figure 3.4 (b) shows that of a channel passivated (CHP) a-Si:H
TFT; both of these devices have inverted staggered-electrode structure. As shown by the arrows in Figure 3.4, electrons
injected from the source electrode cross the a-Si:H i-layer, travel the channel formed at the interface between the gate
insulator and the a-Si:H, cross the i-layer again, and reach the drain electrode. Since the channel thickness is estimated
to be several tens of nanometers, the interface properties play a critical role in determining the TFT characteristics. In
the BCE TFT, the channel length (L) is determined by the design rule or the minimum size (S) of fabrication (L=S). In
the CHP TFT, however, the channel length becomes large: L=S+2/_\L where /_\L is the process margin. A process
margin of 2 um for a design rule of 10 um yields L=14 um. When the drain voltage becomes large (V d >V g ) in the
CHP TFT, electrons flow in two ways at the drain: the bottom and the top channel. The voltage applied to the drain
electrode forms the top channel as shown in Figure 3.4 (b). This effectively reduces the channel length to L=S+(1+k)
/_\L, where k depends on the drain voltage and 0<k<1.

Figure 3.5 Cross-section of a staggered-electrode or top-gate structure a-Si:H TFT. The bold black regions on top of
source and drain electrodes are ohmic contact layers of n+ a-Si:H.

The staggered-electrode structure is shown in Figure 3.5 In this type of TFT, the source and drain electrode are formed
first, and the gate electrode comes at the top of the layers. The channel length is equal to the minimum fabrication size,
S. Not shown in this figure is the light shield layer to block the illumination from the back.

3.2. TFT Characteristics


3.2.1. Current-Voltage Characteristics
The basic current-voltage characteristics of a TFT can be analyzed in essentially the same way as those of their
crystalline silicon counterparts 9 . The following assumptions are made for the formulation: (1) the carrier mobility in
the channel is constant, (2) the gate capacitance is constant and independent of the gate voltage, (3) the source and
drain electrodes are ohmic contacts to the semiconductor, (4) the initial charge density in the semiconductor is n 0 , and
(5) the gradual channel approximation can be applied. The last condition means that the transverse field in the channel
is greater than the longitudinal field (Ex ).

The coplanar structure shown in Figure 3.6 is used for the analysis. The TFT electrode configuration described in
Section 3.1 differs from this structure in that the carriers cross the a-Si:H thin film before reaching the channel.
However, if the a-Si:H film is thin, the staggered-electrode configuration can be analyzed in terms of the coplanar
TFT structure. The effects of a thicker film will be discussed later in

Figure 3.6 The coplanar TFT structure used in the analysis. In a-Si:H TFTs, the staggered electrode configuration is
generally adopted. Therefore, the analytical model is different from the actual device. However the current-voltage
characteristics of such devices can be described well by the analytical results obtained using this model.

Section 3.5. The coplanar structure is not common in a-Si:H TFTs, and the technologies essential for fabricating
coplanar TFTs (such as impurity diffusion or ion implantation) are not well established in a-Si:H.

The application of a gate voltage Vg induces charge density /_\n(x) in the channel region. This is given by:

(3.1)

where C i is the gate capacitance per unit are (= i/d), t is the a-Si:H thickness, d is the gate insulator thickness, and
V(x) is the drain voltage at distance x from the source. If the thickness, t, is assumed to be sufficiently small, the drain
current I d is given by
(3.2)

where 6 0 and /_\6(x) are the initial conductivity and the incremental conductivity due to /_\n(x), respectively. From
Equations (3.1) and (3.2), I d is given by

(3.3)

which reduces to

(3.4)

Then, the drain current is given by

(3.5)

where V t -etn 0 /C i . The threshold voltage V t depends on the initial charge density, n 0 .

Equation (3.5) is valid for a voltage range of 0 V d V g -V t. Beyond this range, the current is assumed to be
constant as in insulated-gate field-effect transistors. Low V d values correspond to the region of linear output
characteristics where the drain conductance, g d , and the transconductance, g m , are given by

(3.6)

(3.7)

The drain conductance is a linear function of V g , and the transconductance is proportional to V d . Saturation of the
drain current occurs when dI d /dV d =0, due to pinch-off of the conducting channel in the neighborhood of the drain. In
this case, the saturation current, I dsat, is given by (square-law relationship)

(3.8)

for V d V g -V t . The transconductance in the saturation region is given by

(3.9)

The high-frequency performance of a-Si:H TFTs can be estimated from the gain-bandwidth product, which is
equivalent to the maximum operating frequency defined for crystalline silicon devices.
(3.10)

Substitution of Equation (3.7) in to Equation (3.10) yields

(3.11)

for V d <V g -V t, and

(3.12)

at the saturation region (V d >V g -V t).

The experimental transfer and output characteristics of a-Si:H TFTs were shown in Figures 3.1 and 3.2. The mobility
un of the a-Si:H TFT shown in Figure 3.1 was un =0.4 cm2 /Vs, but a higher un of about 1.0 cm2 /Vs is obtained in
TFTs as shown in Figure 3.7. The transfer characteristics of Figure 3.7 feature low off-current, low threshold voltage,
steep subthreshold slope, and wide dynamic range, i.e., a high on-current and low off-current. The threshold voltage of
the TFT shown here is about 1 V, and there is virtually no p-type conduction at negative gate bias voltage, because of
the blocking contact at the source and the low hole mobility of a-Si:H. The subthreshold slope is 0.3 V/decade, and the
on/off current ratio of I d is large than 108 .

Two important parameters of TFTs are the carrier mobility u n and the threshold voltage V t. In a-Si:H TFTs, the main
carriers are electrons whose room-temperature mobility is typically 0.3–0.6 cm 2 /Vs. Although higher mobility values
approaching u n =1 cm2 /Vs have been reported, these have only been achieved in carefully prepared samples. The
temperature dependence of the mobility and the threshold voltage is shown in Figure 3.8. The dependence of the
mobility is given by

(3.13)

where u 0 is the extended state electron mobility, N C is the density of states at the mobility edge, n is the total electron
density, and E a is the activation energy, which reflects the tail state distribution of the a-Si:H. In Figure 3.8, E a=0.13
eV, and values from E a=0.07 eV10–0.16 eV11 have been reported.
Figure 3.7 The transfer characteristics of channel-passivated a-Si:H TFTs with a mobility of 1.0 cm 2 /Vs for a drain
voltage of 10 V. The threshold voltage is 1 V.

3.2.2. Parasitic Effects

As described in Section 2.3.2, the parasitic capacitance of the TFT induces a dc component in the ac voltage applied to
the nodes or the pixel electrodes of a TFT/LCD. Since the potential at the node is alternating, the parasitic capacitance
between the source and the gate, C gs, and that between the drain and the gate, C gd, in the static TFT operation both
become important.
Figure 3.8 The temperature dependence of the electron drift mobility and the threshold voltage of the a-Si:H TFT.

In the simple model, the parasitic capacitance is determined by the overlap of the electrode and the source (or drain)
electrode. Therefore, the area used to calculate the capacitance is given by the channel width multiplied by the overlap
length. However, the effective overlap length is longer than the geometrical length due to the fringe effect which has
been shown to correspond to an extra length of about 1 um by extrapolating the linear relation between C gs and the
overlap length. The parasitic capacitance depends on the gate voltage. When the gate voltage is low, the contribution
from the intrinsic amorphous silicon layer must be added to that of the gate insulator. When the gate voltage is turned
on, however, an accumulation layer is formed at the interface between a-Si:H and SiN. This reduces the effective
thickness of the parasitic capacitance to the thickness of the SiN layer in its limit, causing the capacitance to increase.
Figure 3.9 shows how the parasitic capacitance depends on the gate and drain voltages 12 . The dependence of C gs
(Figure 3.9 (a)) shows a rather simple behavior, while that of C gd depends on both V g and V d , as shown in Figure 3.9
(b). In a TFT/LCD panel, we have to consider both of these cases. When the data voltage is positive and the potential
of the node to be charged is negative, the parasitic capacitance behaves as shown in Figure 3.9 (a). In the reverse case
where the data voltage is negative and the node potential is positive, the parasitic capacitance behaves as shown in
Figure 3.9 (b). The variation of capacitance as a function of the drain voltage is shown in Figure 3.10 13 , and the
frequency dependence of C gs is
Figure 3.9 Parasitic capacitance of a BCE type a-Si:H TFT as functions of the gate and the drain voltage 12 .

shown in Figure 3.11 13 . Since the voltage offset is caused by a transient phenomenon, its frequency dependence is
important and adds further complexity to this effect.

The photoconductivity of a-Si:H has been widely studied for applications such as solar cells 14 , photosensors 15 , and
electronic copying machines 16 . Figure 3.12 shows the spectral response of a-Si:H pin diodes 15 . The spectral
response of a-Si:H covers the entire visible spectrum. In TFT/LCD applications, however, the photoconductivity of a-
Si:H is not at all desirable. Since TFT/LCDs are operated under strong illumination, this effect must be minimized to
maintain normal TFT operation. In LCD modules, illumination comes both from the bottom and the top. The backlight
illumination is more intense than the ambient illumination
Figure 3.10 Parasitic capacitance as a function of the drain voltage 13 . Negative drain voltage corresponds to
exchanged drain/source electrodes. The TFT is BCE type in which W/L=72 um/8 um, the overlap length between the
gate and the source (drain) is 5 um, the a-Si:H thickness is 300 nm, and Vt is 4.5 V.

Figure 3.11 Parasitic capacitance vs. frequency 13 . The dependence of capacitance on the gate voltage gets smaller as
the frequency increases. The TFT is the same as that of Figure 3.10.

which comes from the top. Therefore, the a-Si:H island is designed to be smaller than the gate electrode. Since the
gate electrode acts as a light shield, the TFTs do not suffer from leaky photoconductive current.
Figure 3.12 Spectral response of an a-Si:H pin diode illuminated from the p+ a-Si:H layer covered with a transparent
electrode of ITO (Indium Tin Oxide) 15 . The thicknesses of the p, i, and n layers are 20, 550, and 30 nm, respectively.

The photocurrent I ph is given by

(3.14)

where a is the absorption coefficient of a-Si:H and t is the thickness of the film. Naturally, a thinner a-Si:H film is
preferable in order to reduce the photocurrent in the reverse biased TFT. The relation between the a-Si:H thickness and
the photocurrent of the drain current of the TFT due to photocarriers is shown in Figure 3.13 for the case where the a-
Si:H layer is illuminated directly from the top of the TFT 17 . A non-linear dependence on film thickness is observed,
and this dependence is given by

(3.15)

Since a linear dependence is expected from Equation (3.14), other factors such as the hole blocking contact or n+ a-
Si:H layer, must be taken into account to explain this dependence.
Figure 3.13 Photocurrent of the TFT (BCE) vs. the thickness of the a-Si:H layer. An illumination of 5,000 lx
corresponds to direct backlight illumination, which usually comes from beneath.

3.3. Threshold Voltage Shift


The transfer characteristics of a-Si:H TFTs are unstable against the voltage applied to the gate electrode. As shown in
Figure 3.14, this instability manifests itself as a parallel shift of the I-V characteristics to the right when the gate bias
voltage is positive and to the left when the
Figure 3.14 The a-Si:H TFT transfer characteristics shift according to the positive or negative stress voltage applied to
the gate 18 . Along with this threshold voltage shift, a decrease of drift mobility is usually observed.

gate bias is negative. This causes the threshold voltage (V t) to shift. Figure 3.15 shows an example of this shift (/_\V
t) as a function of gate bias voltage. The V t shift can be approximated as ±|V g | B where B varies from 1 to 4
depending on the gate voltage 18 . /_\V t depends strongly on the gate bias, while it varies only slightly with the drain
voltage. The data in Figure 3.15 was obtain under stress conditions in which the source and drain electrodes were
grounded. The annealing process is usually effective in recovering the initial I-V characteristics. Therefore, the same
TFT can be used repeatedly to obtain the /_\V t data. After heat treatment at 200°C for 30–60 minutes, the transfer
characteristics return to their initial curves.

The experimental formulation of /_\V t as a function of temperature T, time, t, and gate voltage, V g , is given by:

(3.16)
Figure 3.15 Dependence of the threshold voltage shift on the stress voltage applied to the gate 18 . Stronger
dependence is observed for negative voltage stress.

where E a is the activation energy, and A, B, and Y are constants. Figure 3.16 shows the dependence of /_\V t on the
stress time. The constant, B, depends more on the bias voltage polarity than the other constants. For a positive bias, B
is in the range of 1–2, and for a negative bias, B becomes rather high ranging from 3 to 4. The constant Y varies from
0.3 to 0.4 depending on the device parameters and the bias polarity. The activation energy E a is in the range of 0.2–0.3
eV. 10 , 18

The constants in Equation (3.16) depend on the deposition conditions of the a-Si:H and gate insulator. The gas mixing
ratio, the rf power, the gas pressure, and the substrate temperature are some of the conditions to be taken into account.
The effects of these conditions on /_\V t are
Figure 3.16 Threshold voltage shift as a function of stress time. The dependence is almost the same for both positive
and negative voltage stress: /_\V t t 0.33 for Vg =20 V and /_\V t t 0.34 for Vg=-10 v.

mixed together and there are no universally applicable guidelines for determining the deposition conditions. The
relation between the SiN deposition conditions and /_\V t has been studied since the internal stress of SiN film is high
(~109 dyne/cm 2 ) and the stress can be both tensile and compressive (Figure 3.22). The stress-free condition does not
necessarily correspond to the lowest V t shift. The stoichiometric composition (Si3 N4 ) and a substrate temperature of
about 300°C are preferred.

The V t shift mechanism has been discussed by Powell 10 , who implied from a simple analogy to conventional MOS
FETs that the charge trapping in the SiN gate insulator causes this shift. This is supported by the fact that TFTs with
silicon-rich SiN are less stable than those with stoichiometric SiN. Another proposed mechanism is that /_\V t arises
due to the formation of bias-induced deep trap levels in the amorphous silicon film. Since the channel is formed in the
vicinity of the a-Si:H/SiN interface, these trap levels may also be associated with the interface. In any case, both
mechanisms are believed to play roles in the appearance of /_\V t. At higher gate bias, the /_\V t of a-Si:H TFTs is due
to charge trapping in SiN, mainly through tunneling. The dominant mechanism at lower gate bias, on the other hand, is
the creation of states or the breaking of bonds in the a-Si:H film at or near the interface.

The V t shift has been studied by illuminating a sample with an external light 19 , 20 . The increase of /_\V t with
increasing illumination reminds us of the defect creation kinetics in bulk a-Si:H or the Staebler-Wronski effect 21 .
The dependence of the defect creation on the light flux is given by 22

(3.17)

where N D is the defect density, G is the illumination intensity, and t is time. The dependence of defect creation on
time in Equation (3.17) is similar to the time dependence of /_\V t shown in Figure 3.16. However, the carriers
generated by the illumination are not directly reflected in /_\V t, and only a small V t shift is observed as shown in
Figure 3.17 20 . The photoexcited carriers are distributed in the a-Si:H layer and do not accumulate at the a-Si:H/SiN
interface. Charge accumulation does not occur without a bias voltage, as is seen from the energy diagram of the a-Si:H
and SiN system shown in Figure 3.18. However, when the gate is biased, the photocarriers accumulate at the a-
Si:H/SiN interface resulting in a large /_\V t in the illuminated TFT than in a TFT that is not illuminated.

Instability in TFT operation and the threshold voltage behavior have been studied over a timespan of from 103 seconds
to one month
Figure 3.17 Threshold voltage shift of an a-Si:H TFT. The illumination does not directly result in the V t shift.
However, the application of a gate bias increases the shift. The TFT was illuminated by a 1,000 lx light and V g =-30V
and Vd =Vs =0.

Figure 3.18 Energy band diagram of the a-Si:H/SiN system. The band bending at the interface corresponds to the flat
band voltage of about -2 V.

(2.6×10 6 s). The lifespan can be estimated by extrapolating the /_\V t behavior in this period, but it is difficult to
predict exactly the V t shift that would occur over a period of years from the data of just one month. Moreover, the
TFTs used in active-matrix liquid-crystal displays are operated in pulsed mode and dual polarity mode. In order to
estimate the lifetime of TFTs in a TFT/LCD panel, we assume that the additive law can be applied, i.e., the total
threshold voltage shift is the sum of the V t shifts due to the positive bias and negative bias. However, TFTs behave in
a complex manner when operated in pulsed mode as described above, and it is not very easy to evaluate their real
lifetime. In practice, the lifespan of TFTs in LCD panels seems to be longer than that estimated from the dc V t shift
data. Figure 3.19 shows an example of the long life-time that can be achieved in LCD panel TFTs. In this example, the
LCD/TFT panel was an assumed VGA monitor operated at a frame rate of 60 Hz. In this figure, /_\V t clearly
saturates, indicating that the real lifetime is longer than that predicted with the dc data (e.g. Figure 3.16).

Figure 3.19 Threshold voltage of an a-Si:H TFT as a function of the operating time. The gate was biased to -20 V to
compensate for the pulsed bias of 20 V. The pulse width and the repetition period were fixed at 33 us and 16.6 ms,
respectively.

This is due to a number of reasons: (i) the tendency for positive and negative shifts to cancel each other out, (ii) the
frequency dependence of /_\V t, and (iii) the dependence of /_\V t on the on/off time ratio.

The dependence of /_\V t on the frequency of the voltage pulse is shown in Figure 3.20. /_\V t changes little with
frequency when the TFT is operated with positive bias pulses. However, when the bias voltage is negative, /_\V t
decreases as the frequency of the voltage pulses increases. Moreover, /_\V t has been found to depend on the duty ratio
of the pulse voltage as shown in Figure 3.21. The total stress time was kept constant in the experiment. It is worth
noting that there is a jump in /_\V t at negative bias as the duty ratio changes from 100 to 99%. /_\V t is smaller in
pulsed mode than in the dc case 23 . These results can be explained by an equivalent circuit model. When the gate
voltage is applied to the TFT in a pulsed or ac mode, the voltage is divided by the RC circuit of the SiN layer and that
of the a-Si:H layer. Consequently, the effective voltage applied to the gate insulator or the electric field at the interface
of SiN/a-Si:H depends on the frequency of the ac pulse.
Figure 3.20 Threshold voltage shift as a function of the frequency of the pulse voltage.

Figure 3.21 Dependence of the threshold voltage shift on the duty ratio of the pulsed voltage applied to the gate 23 .
The frequency of the pulse is 60 Hz. The total stress time was kept constant for all the data points.

3.4. Process-Related Issues


This section describes the process-related issues of a-Si:H TFTs. Back-channel-etched (BCE) TFTS as shown in
Figure 3.4 (a), are fabricated as follows: the Cr is first deposited at a typical substrate temperature of 150°C and an
argon gas pressure of ~1 mTorr. It is then photolithographically patterned to form the gate electrode by wet etching
with an etchant of nitric acid cerium ammonium. A p-CVD system is used to form three layers of amorphous material:
silicon nitride, intrinsic amorphous silicon, and n-type (phosphorus doped) amorphous silicon. The n-type a-Si:H acts
as an ohmic contact layer for the intrinsic amorphous silicon. To deposit the intrinsic a-Si:H layer, silane (SiH 4 ) gas
diluted with hydrogen is introduced into the reaction chamber while the substrate is heated to between 250° and 350°
depending on the film specifications. The gas pressure is usually in the range of 0.5-1 Torr and the gas is decomposed
to form the a-Si:H layer in an rf (13.56 MHz) excited plasma. To deposit the n-type a-Si:H layer, phosphine (PH3 ) gas
diluted with hydrogen is introduced in addition to silane gas.

To deposit silicon nitride, silane, nitrogen and ammonia are mixed and introduced into the CVD chamber. The ratio of
gas flow is determined according to the resultant morphology, the internal stress of the film, the insulating properties
and the characteristics of the interface between SiN and a-Si:H. These properties also depend on the temperature,
pressure, and the rf power of plasma excitation. Figure 3.22 shows an example: the dependence of the SiN film stress
on the rf power 24 . This film was deposited under a mixture of ammonia, silane and hydrogen. Dry etching with SF6
or CF4 gas is used to etch both a-Si:H and SiN.

Next, Cr and Al are sputter-deposited and etched to form the source and drain electrodes. The Cr is used as a buffer
metal to prevent the Al from diffusing into the a-Si:H, while the aluminum is necessary to reduce the resistance of the
data lines in the TFT panel. A thin layer of Cr (<10 nm) provides enough buffering action. The next step is to etch off
the n-type a-Si:H layer. The source and drain electrodes that are already patterned act as a mask for etching the n-type
a-Si:H. The final step is the formation of an insulating SiN layer to passivate the exposed surface of intrinsic
amorphous silicon.

The fabrication of CHP TFTs (Figure 3.4 (b)) differs from that of BCE TFTs in the SiN/a-Si:H/SiN layers are
deposited instead of

Figure 3.22 Dependence of the internal stress of SiN film on the rf power applied to the capacitive electrodes 24 . The
deposition conditions are: SiH4 = 15 sccm, NH 3 =90 sccm, H2 =200 sccm, cathode area=900 cm2 , pressure = 1 Torr,
and Tsub =300°C.

SiN/a-Si:H/(n + ) a-Si:H layers. The SiN channel passivation layer is patterned photolithographically to define the
channel length. Then, a thin layer of n+ a-Si:H is deposited and an amorphous silicon island is defined. The source and
drain electrodes are formed as described above, and the n+ layer on top of the passivation layer is etched off using
these electrodes as a mask.

In the CHP TFT, the etching of the n+ a-Si:H layer stops automatically when the SiN surface is exposed. Hence, the
control of the etching time is not so critical. However, in the BCE TFT, the etching time has to be long enough to
completely etch the n+ layer down to the intrinsic layer of amorphous silicon. As a result, the i-layer cannot be made
as thin as in the CHP TFT. The advantage of CHP TFTs is that the leakage current due to a-Si:H photoconductivity
can be decreased by making the intrinsic layer thin. Uniformity in device characteristics is another feature, but the
parasitic capacitance becomes relatively large due to the thin a-Si:H layer. The merit of BCE TFTs is that they require
fewer process steps than CHP TFTs.

Many types of gate metals and gate insulators have been used in TFT fabrication. Examples of gate metals include Cr
25 , Ta 26 , Mo 27 , Mo-Ta 28 , and Al 29 , and examples of gate insulators include SiN, SiOx , Ta2 O 5 26 , MoTaOx 28 ,
and Al2 O3 29 , and combinations of these materials. These insulators are deposited by p-CVD (SiN and SiO x ,
sputtering (SiO x and TaO x ) or anodic oxidation (Ta 2 O5 and Al2 O3 ). Stacked-layer structures like Ta2 O5 /SiN and
Al2 O3 /SiN are especially important for the gate insulator. In these structures, the Ta or Al gate metal is oxidized by
anodic oxidation to form Ta2 O5 or Al2 O3 , and SiN is then deposited on these oxidized layers. The SiN layer is
necessary to secure good interface quality between a-Si:H and the gate insulator as well as to obtain good insulating
characteristics. The stacked-layer system is preferred, since it is formed by a combination of wet and dry processes.
This combination contributes to reducing the defect density of the overall gate insulator system, and leads to good step
coverage of the metal by the oxide film.

Figure 3.23 shows the current-voltage characteristics of various insulators. The leakage current level of Al2 O3 is far
lower than that of Ta2 O5 and comparable to that of SiN. Its breakdown field of ~7×106 V/cm is also comparable to that
of SiN. The I-V characteristics of Ta2 O5 sandwiched between metals (Metal-Insulator-Metal) is given by 30

(3.18)

where a and B are coefficients related to the materials and geometry. This is the Poole-Frenkel current which is
expressed as

(3.19)
Figure 3.23 Current-voltage characteristics of various insulators 29 . All samples have areas of 1 mm2 . The breakdown
voltage of the Al2 O3 film is ~150 V corresponding to a field of ~7×106 V/cm.

where J is the current density, E is the electric field in the insulator, is the potential barrier at the contact, e is the
electronic charge, i is the permittivity of the insulator, k is the Boltzmann constant, and T is the absolute temperature.
The MIM diodes made of Ta2 O5 is also used to address the LCD panels 30 .

3.4.1. Aluminum Gate TFT

As discussed in Section 2.3.3, it is essential to use low-resistivity metal for the gate metallization to reduce the gate-
delay of the TFT/LCD

panel. Aluminum is a low-resistivity metal as shown in Table 3.1. It is widely used for LSI metallization, and is a
natural choice for the gate electrodes. The use of anodic oxidation technology is the key to developing aluminum gate
TFTs as shown in Figure 3.24 29 . The Al2 O3 is formed by anodically oxidizing the Al, forming a dense film with
insulating characteristics comparable to those of silicon nitride.

Another feature of Al2 O3 is that it suppresses the formation of hillocks and whiskers. Aluminum tends to develop
hillocks and whiskers

Figure 3.24 Cross-section of aluminum gate a-Si:H TFT. The gate insulator consists of silicon nitride and aluminum
oxide. The relative dielectric constant of Al2 O3 (=9.2) is higher than that of SiN (= 6.9), and high transconductance
can be expected.

when it is heat treated or covered with a thin film. For example, if a silicon nitride film is directly deposited on
aluminum, the surface of the metal becomes rough due to the formation of many hillocks. High internal stress in the
SiN film escalates this formation process. However, if the SiN film is deposited on anodically oxidized aluminum, no
hillocks develop and the SiN surface remains smooth. The aluminum oxide also helps suppress defects—the defect
density of the stacked-layer SiN/Al 2 O3 system on Al was found to be several times lower than that of the single-layer
system of SiN on a chromium electrode. The low defect density of the gate insulator contributes directly to lowering
the incidence of malfunctioning TFTs, leading to higher TFT/LCD yield.

Figure 3.25 (a) shows the principle of anodic oxidation. The Al is connected to the anode of a dc power supply and
platinum is connected to the cathode. These electrodes are immersed in a solution and the Al is oxidized in constant
current mode. In this mode, the voltage across the oxide film increases linearly as shown in Figure 3.25 (b). The
electrochemical reaction proceeds as follows

Figure 3.25 (a) Anodic oxidation system and (b) its operation mode showing the voltage and current variation with
time. The eighty-percent rule applies to the oxidized film, i.e., the breakdown voltage of the film is approximately
equal to the final oxidation voltage multiplied by 0.8.

2Al+3H2 O Al2 O3 +6H+ +6e-

The solution consists of ethylene glycol and tartaric acid. After a fixed interval, the oxidization is switched to constant
voltage mode, so that the current decays rapidly and approaches zero. The thickness of Al2 O3 is determined by the
voltage applied in constant voltage mode. The proportionality factor is 1.4 nm/V. The oxidation time is inversely
proportional to the current during constant current mode. Since the thickness of aluminum oxide depends solely on the
maximum voltage, uniformity of the film thickness over the area is better than ±0.5%. This contributes greatly to the
reproducibility and uniformity of the TFT/ LCD panel fabrication. In a dry process like CVD, uniformity is
approximately one order of magnitude worse than this.

3.4.2. Self-Alignment Process

The self-alignment process of TFTs makes use of the fact that the glass substrate is transparent. Self-alignment is
achieved by shining light from below the substrate. In general, self-alignment is important for reducing the size of
devices, and this also applies to TFT fabrication where smaller device size and reduced process margins contribute to
achieving a larger pixel aperture ratio in TFT/LCD panels. However, the most important aspect of self-alignment in
TFT fabrication is reduction of the parasitic capacitance.

In self-aligned TFT fabrication, the devices are illuminated by UV light both from below and from above the
substrate. When light is shone from below the glass substrate, the patterned gate electrode acts as a mask, and when
light is shone from above the glass, a normal photolithographic mask is used. A small portion of the light shone from
below is absorbed in the a-Si:H layer. However, in CHP TFTs where self-alignment technology is important due to
the large parasitic capacitance, the amorphous silicon layer is generally thin and UV light absorption is not a serious
problem. The self-alignment process for a-Si:H TFTs is shown in Figure 3.26 31 . The lift-off technique is used to
fabricate the TFT, and there is little overlap between the gate and source/drain electrodes. In order to achieve good
ohmic contacts, an overlap length of 1-2 um is necessary between the electrodes. In case of no overlap (no gap), the
drain current decreases by about 10% from the full injection. When there is a gap of 1 um between the electrodes, the
drain current drops further down below
Figure 3.26 Steps of the self-alignment process 31 . The lift-off technology shown here is not always needed to
fabricate self-aligned TFTs.

10% of the full overlap. Therefore, in the self-alignment process, it is important to produce this overlap accurately,
uniformly, and reproducibly from run to run. The length of this overlap depends on the doping level of the ohmic
contact layer. If the doping level is increased, the overlap length can be reduced. For this purpose, several schemes
have been proposed, e.g., the use of ion-implantation technology 32 and microcrystalline silicon technology.

3.5. Computer Simulation of a-Si:H TFT


The analytical model described in Section 3.2 provides a good description of the electrical characteristics of a-Si:H
TFTs. However, in order to simulate TFT behaviors more precisely, it is necessary to include the effect of electrons
trapped in localized states in the energy gap and the effect of carriers crossing the intrinsic amorphous layer of a-Si:H
in the staggered electrode TFT structure. An improved version of the crystalline silicon model incorporating these
effects was simulated in a two-dimensional a-Si:H TFT simulator 33 based on a general-purpose three-dimensional
crystalline silicon device simulator.

The device structure used in the simulation is shown in Figure 3.27. The total electron density, n, in the a-Si:H region
consists of the electron density n c in the conduction band and the trapped electron density n T (n =n c +n T ). The hole
density is disregarded because the n+ a-Si:H region blocks hole injection and the hole mobility is low. It is assumed
that the energy distribution of these electrons is given by the quasi-thermal equilibrium condition caused by frequent
electron transitions between the extended states and the localized states in the gap. The quasi-Fermi energy is given by
the Boltzman distribution function as

Figure 3.27 TFT structure used in the device simulation. The source and drain electrodes are not shown here since the
series resistance of n+ a-Si:H is low enough not to affect the device characteristics. We assume the gate is made of
chromium and the passivation layer is SiN.

(3.20)

where N c is the effective state density at the conduction band edge (mobility edge), k is the Boltzman constant, T is
the absolute temperature, E c is the conduction band edge energy, and E F is the Fermi energy.

The trapped electron density n T is given by the Fermi-Dirac distribution as

(3.21)

where E v is the valence band edge energy, and N(E) is the localized state density in the energy gap. From Equations
(3.20) and (3.21), n c and n T are functions of E F , and therefore n T can be regarded as a function of n C ,

(3.22)

In the a-Si:H region, the Poisson equation is given by

(3.23)

where is the electrostatic potential, s is the permittivity of a-Si:H, e is the electronic charge, and N D is the
effective donor density. The current continuity equation is given by

(3.24)
(3.25)

where J n is the electron current density and u n is the electron mobility.

The potential in the insulator is described by the Laplace equation, and Gauss's law holds for the internal interface
boundary. The Dirichlet condition is applicable to the surface electrodes assuming thermal equilibrium and charge
neutrality, and the Neumann condition is applicable to the other outer surfaces. Equations (3.23), (3.24) and (3.25) are
the basic equations for the steady-state analysis of a-Si:H TFTs. The finite difference method was used to convert these
equations to discrete form, and Newton's iterative method was used to solve them on a supercomputer.

The sample was assumed to have an a-Si:H thickness, t, of 0.3 um, gate insulator (SiN) thickness, d, of 0.3 um, gate
length of 20 um and channel length (source-drain spacing) of 13 um. The overlap between

Figure 3.28 Localized state density distribution obtained from the quasi-static C-V measurements (voltage scan rate:
0.02 V/s) of an a-Si:H/SiN/metal capacitor 33 . The origin of the energy was taken to be the equilibrium Fermi energy
of a-Si:H. Ec and Ep denote the conduction band edge and the subpeak energy of the gap states.

the gate and the source electrode (between the gate and the drain electrode) is 3.5 um. The effective donor density N d
of the n+ layer was assumed to be 1021 cm-3 , which is high enough to avoid the effect of series resistance. The
specific dielectric constants of a-Si:H and SiN were assumed to be 11.7 and 7, respectively.

The localized state density distribution N(E) obtained from C-V measurement of an a-Si:H/SiN/metal capacitor is
shown in Figure 3.28 (the flat band voltage is assumed to be V FB=-2V). This function is expressed as a linear
combination of exponential functions as follows:
(3.26)

(3.27)

where E C is the conduction band edge energy, E P is the subpeak shown in Figure 3.28, D, D 1 and D 2 are the
characteristic energies, and N c and N P are the characteristic densities of states. Data matching gives us the values of
these parameters: D=24.7 meV, D 1 =50 meV, D 2 =56 meV, N c=1.7×10 21 cm- 3 eV-1 , N P =1.9×10 17 cm-3 eV-1 , and E
C -E P = 0.46 eV. The detailed structure of the state density at energies far below E p peak is disregarded because its
effects on the device characteristics are negligible. For the localized state density distribution given by Equations (3.26)
and (3.27), the relation between n T and n C has been calculated as shown in Figure 3.29. For n C larger than 1016
cm- 3 , n C is nearly proportional to n T and about 20 times smaller than n T.

Calculated transfer characteristics for the values listed above are compared with the experimental results in Figure
3.30. The agreement is satisfactory over five orders of magnitude. The electron mobility above the mobility edge was
assumed to be 7 cm2 /Vs. The calculated and experimental output characteristics are shown in Figure 3.31.

Figure 3.29 Relation between trapped electron density n T and mobile electron density n C. The localized state density
is shown in Figure 3.28 and the effective state density at the conduction band edge was assumed to be 1.7×10 21 cm-3
eV-1 .
Figure 3.30 Comparison between the calculated and experimental transfer characteristics of an a-Si:H TFT 33 . The
channel width is 455 um.

Again, the agreement is good enough to simulate even the slight crowding characteristics observed at low drain
voltages. The crowding characteristics become more prominent as the a-Si:H layer gets thicker due to the increase in
series resistance. The simulated dependence of the threshold voltage, V t, on temperature also agrees with the
experimental results as shown in Figure 3.32. The threshold voltage is defined as the extrapolation of the vs. Vg
characteristics.

This simulator provides further information on the a-Si:H TFT characteristics. One example is the C-V characteristics.
The capacitative behavior between the gate and source is quite different from that between the gate and drain. In order
to estimate the gate capacitance, the gate charges Q g are calculated as follows (Gauss's law):

(3.28)
Figure 3.31 Comparison between the calculated and experimental output characteristics 33 . The a-Si:H layer is rather
thick (0.3 um), and slight crowding characteristics are seen at low drain voltages. When the output characteristics are
calculated with an a-Si:H thickness of 0.4 um, the crowding characteristics become more prominent which agrees well
with the experimental results.

Figure 3.32 Temperature dependence of the threshold voltage of a-Si:H TFT. The calculated results agree well with
the experimental results.

where E n is the normal component of the electric field on the gate electrode surface and integration is carried out over
the gate electrode surface. An increment of Q g (i.e., /_\Q g ) divided by an increment of /_\V d (/_\V s ) gives the drain-
gate (source-gate) capacitance. The results are shown in Figures 3.33 and 3.34. In the C gs curve of Figure 3.33, a
hump appears around V g =0 V. This turns out to be related to the subpeak in the localized state density in Figure 3.28.
If a subpeak density of one-tenth the measured value (1.9×10 16 cm- 3 eV- 1 ), is substituted into the calculation, the
hump in the C-V characteristics disappears. The original C-V characteristics of the a-Si:H /SiN / metal capacitor that
gave the localized state density of Figure 3.28 also have a structure with a less noticeable hump than that of Figure
3.33.
The difference between a-Si:H TFTs and MOS FETs is that the carriers (electrons) cross the semiconductor layer at the
source and drain electrodes. The simulation results show that the current flow is well confined

Figure 3.33 Gate-source and gate-drain capacitance as functions of the gate voltage. The device structure used in this
simulation is shown in the insert. This structure is closer to the real device than that of Figure 3.27. The C-V
characteristics are sensitive to the cross-sectional shapes of the layers. The drain voltage is 5 V.

Figure 3.34 Gate-source and gate-drain capacitance as functions of the gate voltage. The difference between this
figure and Figure 3.33 lies in the subpeak density NP of Figure 3.28, which is lowered to one-tenth of the measured
value. (NP =1.9×10 16 cm-3 /eV -1 ). The hump seen in Figure 3.33 disappears almost completely.

within the "source path" and the "drain path" (Figure 3.35) which are the paths of carriers crossing the a-Si:H layer,
and that the potential distribution in these paths closely resembles that of the space-charge-limited currents observed in
vacuum tubes or in high-resistance semiconductors.

The space-charge-limited current is derived as follows. We suppose that a voltage, V, is applied across the thickness, t,
of an a-Si:H layer. Then, the electric field, E, across the a-Si:H layer is described by Poisson's equation:

(3.29)

where the space charge was approximated by trapped electrons and r 1 is the proportionality factor of n T/n C. The
current density J is given as

(3.30)

neglecting the diffusion current. Considering the current density to be constant with respect to y, we obtain

Figure 3.35 Cross section of an a-Si:H TFT. The carriers injected from the source cross the a-Si:H layer by way of the
source and drain paths.

(3.31)

Integrating Equation (3.31), we get

(3.32)

The currents in the source path and in the drain path are then given by
(3.33)

(3.34)

where W is the channel width, and L s and L d are the effective overlap lengths in the source path and the drain path,
respectively. V C1 and V C2 are the channel potentials at the opposite ends of the a-Si:H layer with respect to the
source and the drain electrode, respectively.

By using the gradual channel approximation, the channel current can be expressed as

(3.35)

where i is the permittivity of SiN, V t is the threshold voltage, L is the channel length, and r 0 is the approximate
ratio of n T /n C in the channel, which is different from that in the source and drain paths, that is, r 1 . From Equations
(3.33), (3.34) and (3.35), and assuming L s =L d (=L ov), the following equation for the drain current is obtained:

(3.36)

where

This analytical model is a good approximation to the rigorous calculations performed by a computer. The effect of the
a-Si:H thickness is condensed in the expression of V0 in Equation (3.36). It modifies the TFT characteristics, causing
them to depart from the ideal co-planar TFT model. One example is the crowding characteristics shown in Figure 3.31.

Another example is shown in Figure 3.36, where the inverse of gain factor 1/B (=L/u n C i W) is plotted as a function
of the channel length,
Figure 3.36 Inverse of the gain factor vs. the channel length 34 . The effective channel length becomes larger than the
channel length due to the source path and drain path of the staggered-electrode TFT.

L (the spacing between source and drain in Figure 3.35). If the relation between the channel length and the gain factor
holds, the extrapolation of the data should coincide with the origin of the graph. However, a clear offset is seen and
this can be explained as the effect of t. Simple analysis leads to the incremental parasitic channel length dL, i.e., the
apparent channel length is given by L+dL. dL is given by 34

(3.37)

where

With values of L=12.4 um, L ov=3.8 um, i/d=18.5 nF/cm2 , s =10.5 o , r 0 =30, and r 1 =180, dL is calculated to be
2.7 um and 8.4 um for a-Si:H thicknesses of 0.5 um and 1.0 um, respectively. These values agree well with the
experimental results.

As the effective channel length decreases and the a-Si:H thickness increases, the effective mobility becomes lower
unless corrections are made. The observed field effect mobility u obs is given by

(3.38)

From Equation (3.37) and (3.38),

(3.39)
The mobility u n in Equation (3.39) corresponds to the drift mobility in the channel, while u obs is the mobility
determined with the inclusion of what we call the parasitic channel length. The mobility calculated from Equation
(3.38) is constant regardless of the channel length. The mobility u obs is lower than u n , and the difference becomes
more prominent as the channel length decreases and the film thickness increases.

References
1. Weimer, P.K. (1961). An evaporated thin-film triode. In Proceedings of the IRE-AIEE Solid State Device Research
Conference (Stanford, California, 1961).

2. Weimer, P.K. (1962). The TFT—a new thin-film transistor. Proc. IRE , 50, 1462–1469.

3. Brody, T.P., Asars, J.A., and Dixon, G.D. (1973). A 6×6 inch 20 lines-per-inch liquid-crystal display panel. IEEE
Trans. Electron Devices , ED-20, 995–1001.

4. Luo, F-C, Hester, W.A., and Brody, T.P. (1978). Alphanumeric and video performance of a 6"×6" 30 lines-per-inch
thin-film transistor liquid-crystal display panel. In Digest of Technical Papers of the Society for Information Display
International Symposium (San Francisco, 1978). California: SID.

5. LeComber, P.G., Spear, W.E., and Gaith, A. (1978). Amorphous-silicon field-effect device and possible application.
Electronics Letters , 15, 179–181.

6. Snell, A.J., Mackenzie, K.D., Spear, W.E., and LeComber, P.G. (1981). Application of amorphous silicon field
effect transistors in addressable liquid-crystal display panels. Applied Physics , 24, 357–362.

7. Spear, W.E., and LeComber, P.G. (1972). Investigation of the localized state distribution in amorphous Si films.
Journal of Non-Crystalline Solids , 8–10, 727–738.

8. Okubo, Y., Nakagiri, T., Osada, Y., Sugata, M., Kitahara, N., and Hatanaka, K. (1982). Large-scale LCDs addressed
by a-Si TFT array. In Digest of Technical Papers of the Society for Information Display International Symposium (San
Diego, 1982). California: SID.

9. Sze, S.M. (1969). Physics of Semiconductor Devices , pp. 567–573. New York: Wiley-Interscience.

10. Powell, M.J. (1983). Charge trapping instabilities in amorphous silicon—silicon nitride thin-film transistors .
Applied Physics Letters , 43, 597–599.

11. Lustig, N. and Kanicki, J. (1989). Gate dielectric and contact effects in hydrogenated amorphous silicon-silicon
nitride thin-film transistors, Journal of Applied Physics , 65, 3951–3955.

12. Ohta, M., Tsumura, M., Ohida, J., Ohwada, J., and Suzuki, K. (1992). Active matrix network simulator considering
non-linear C-V characteristics of TFTs intrinsic capacitances. In Proceedings of the 12th International Display
Research Conference (Hiroshima, 1992), pp. 431–434. California: SID.

13. Nakazato, M., and Higuchi, T. (1992). Capacitance-voltage characteristics of a-Si TFTs. In Proceedings of the
12th International Display Research Conference (Hiroshima, 1992), pp. 439–442 . California: SID.

14. Carlson, D.E. and Wronski, C.R. (1976). Amorphous silicon solar cell. Applied Physics Letters , 28, 671–673.

15. Yamamoto, H., Baji, T., Matsumaru, H., Tanaka, Y, Seki, K., Tsukada, T., et al. (1983). High speed contact type
linear sensor array using a -Si pin diodes. In Extended Abstracts of the 15th Conference on Solid State Devices and
Materials (Tokyo, 1983), pp. 205–208. Tokyo: The Japan Society of Applied Physics.

16. Shimizu, I., Komatsu, T., Saito, K., and Inoue, E. (1980). a-Si thin film as a photo-receptor for electrophotography.
Journal of Non-Crystalline Solids , 35 & 36, 773–778.

17. Sasano, A. Unpublished work.

18. Kaneko, Y., Sasano, A., Tsukada, T., Oritsuki, R., and Suzuki, K. (1986). Improved reliability in amorphous
silicon thin-film transistors. In Extended Abstracts of the 18th International Conference on Solid State Devices and
Materials (Tokyo, 1986). pp.669–702. Tokyo: The Japan Society of Applied Physics.

19. Katayama, M., Morimoto, H., Yasuda, S., Takamatu, T., Tanaka, H., Hijikigawa, M. (1988). High-resolution full-
color LCDs addressed by double-layerd gate-insulator a-Si TFTs . In Digest of Technical Papers of the Society for
Information Display International Symposium (Anaheim, 1988), pp. 310–313. California :SID

20. Kaneko, Y. Unpublished work.

21. Staebler, D.L. and Wronski, C.R. (1977). Reversible conductivity changes in discharge-produced amorphous Si.
Applied Physics Letters , 31, 292–294.

22. Street, R.A. (1991). Hydrogenated Amorphous Silicon , p. 216. Cambridge: Cambridge University Press.

23. Oritsuki, R., Horii, T. Sasano, A., Tsutsui, K., Koizumi, T. Kaneko, Y., and Tsukada, T. (1991). Threshold voltage
shift of a-Si TFTs during pulse operation. In Extended Abstracts of the International Conference on Solid State
Devices and Materials (Yokohama, 1991). pp. 635–637. Tokyo: The Japan Society of Applied Physics.

24. Kobayashi, I., Ogawa, T., and Hotta, S. (1991). Plasma enhanced CVD of silicon nitride film. In Digest of Papers
of the 4th MicroProcess Conference (Kanazawa, 1991), pp. 158–159. Tokyo: Japan Society of Applied Physics.

25. Suzuki, K., Suzuki, H., Tsukada, T., and Kawakami, H. (1985). Amorphous silicon active matrix addressing color
LCD. In Proceedings of the International Display Research Conference , pp. 14–17. California: SID.

26. Ishii, Y, Takafuji, T., Yano, K., Take, H., Funada, F., Matsuura, M., and Wada, T. (1985). High-performance a-Si
TFT array for liquid-crystal display device. In Digest of Technical Papers of the Society for Information Display
International Symposium (Orlando, 1985), pp. 295–296. California: SID.

27. Suzuki, K., Aoki, T., Ikeda, M., Okada, Y, Zohta., and Ide, K. (1983). High-resolution transparent-type a-Si TFT
LCDs. In Digest of Technical Papers of the Society for Information Display International Symposium (Philadelphia,
1983), pp. 146–147. California: SID.

28. Dohjo, M., Aoki, T., Suzuki, K., Ikeda, M., Higuchi, T., and Oana, Y (1988). Low-resistance Mo-Ta gateline
material for large-area a-Si TFT-LCDs. In Digest of Technical Papers of the Society for Information Display
International Symposium (Anaheim, 1988), pp. 330–333. California: SID.

29. Yamamoto, H., Matsumaru, H., Tsutsui, K., Konishi, N., Nakatani, M., Shirahashi, K., Sasano, A., and Tsukada,
T., (1990). A new a-Si TFT with Al2 O3 /SiN double-layered gate insulator for 10.4-inch diagonal multicolor display.
In Technical Digest of the International Electron Devices Meeting (San Francisco, 1990), pp. 851–854. New York:
IEEE.

30. Morozumi, S., Ohta, T., Araki, R., Sonehara, T., Kubota, K., Ono, Y., et al. (1983). A 250×240 element LCD
addressed by lateral MIM. In Digest of Technical Papers of the Society for Information Display International
Symposium (Kobe, 1983), pp. 404–407. California: SID.

31. Asama, K., Kodama, T., Kawai, S., Nasu, Y., and Yanagisawa, S. (1983). A self-alignment processed a-Si TFT
matrix circuit for LCD panels. In Digest of Technical Papers of the Society for Information Display International
Symposium (Philadelphia, 1983), pp. 144–145. California: SID.
32. Akiyama, M., Uchikoga, S., Sakakubo, T., Koizumi, T., and Suzuki, K. (1991). In Digest of Technical Papers of
the Society for Information Display International Symposium (Anaheim, 1991), pp. 10–13. California: SID.

33. Toyabe, T., Masuda, H., Kaneko, Y., Sasano, A., Fukushima, H., and Tsukada, T. (1986). A two-dimensional
numerical model of amorphous silicon thin-film transistors. In Technical Digest of the International Electron Devices
Meeting (Washington, D.C., 1986), pp. 575–578. New York: IEEE.

34. Kaneko, Y., Toyabe, T., and Tsukada, T. (1992). Analysis of effective channel length in amorphous silicon thin-
film transistors. Japanese Journal of Applied Physics , 31, 3506–3510.

CHAPTER 4
Hydrogenated Amorphous Silicon
The electrical conductivity of a-Si:H has been reported to have been controlled over many orders of magnitude by
substitutional doping 1 . Conductivity modulation was observed even with a minute inclusion of phosphorus or boron
into the structure. This aroused a great deal of interest among people in the world of amorphous semiconductor
physics, since it had been believed that the amorphous semiconductor was a dirty material and that substitutional
doping was impossible. The report reminded them of the success of the crystalline semiconductor because
substitutional doping of crystalline semiconductors played a highly significant role in the development of solid-state
electronics and the semiconductor industry.

Since then, there have been many efforts to use this material in various devices and applications. The a-Si:H pin
diodes showed good photovoltaic effects as well as excellent rectifying characteristics and they have been applied to
solar cells 2 . A conversion efficiency of over 10% has been achieved in a-Si:H solar cells 3 . Because of its
photoconductive behavior and the hardness of the material, it is used in copying machines or laser-beam printers in
which the surface of an aluminum drum is covered with an a-Si:H film. The a-Si:H film is resistant against scratches
and fatigue. Also a contact linear sensor array was developed and applied to compact facsimile machines 4 . The a-
Si:H thin-film transistors (TFTs) were developed by the Dundee group who suggested that this TFT could be used in
active-matrix liquid-crystal displays 5 , 6 . They carried out field effect experiments prior to TFT development to
investigate the density of the localized states of a-Si:H 7 .

The success of a-Si:H was brought about by preparing it by decomposing silane in an rf glow discharge or plasma
chemical vapor deposition method. This method was originally developed by Chittick 8 . The silane decomposition
results in the inclusion of hydrogen which compensates for the dangling bonds of the random network structure of
amorphous semiconductors. The inclusion of hydrogen causes hydrogenated amorphous silicon to differ greatly from
evaporated amorphous silicon, although this inclusion was not recognized explicitly in their original works 1 .

In the following sections, the physical aspects of amorphous semiconductors are described in generic form. In later
sections, the physical properties of the hydrogenated amorphous silicon are described in detail.

4.1. Amorphous Semiconductors


The periodic atomic structure of crystalline material leads to Bloch's theory, to account for the band structure and
electronic characteristics. Amorphous semiconductors are distinguished from crystalline semiconductors by their
disordered atomic configurations. However, even if they do not have long-range order they still have short-range
order, i.e., bond length, bond angle, and coordination number similar to those of crystalline semiconductors. In most
amorphous semiconducting materials, the chemical bonding is covalent bonding. The electrons of an isolated atom
occupy s and p bonding states as well as core states which do not contribute to the bonding. Four orbitals (one s and
three p's) can be filled with up to eight electrons with their electron spin up and down from the Pauli exclusion
principle. Figure 4.1 shows the covalent bonding configuration of atoms from group-I to group-VIII of the periodic
table. The group-N elements of the periodic table have N valence electrons. Then, the optimum number of covalent
bonds is 8-N for N 4 and N for N 4. This coordination scheme of atoms is known as the 8-N rule 9 . In an atomic
configuration, two electrons first occupy the s-orbital then other electrons fill the higher-energy p-orbitals. However, in
molecules or in solids with covalent bonding, the total energy is lower in hybrid configurations between s and p states
in which the s-orbital electrons are excited to the p-orbitals.

The silicon atom has four valence electrons which form hybrid molecular orbitals to give four sp 3 orbitals. These
orbitals are composed of one-fourth of an s state and three-fourths of one of the three equivalent p states. The Si atom
is in fourfold coordination. Since Figure 4.1 is a representation of a covalent bonding configuration, it does not apply
to ionic crystals. In ionic bonding crystals like NaCl, one electron moves

Figure 4.1 Electron configuration of covalent bonding elements. The usual configuration is the occupation of s orbitals
first by two electrons, and then occupation of p orbitals. In covalent bonding, however, the hybrid system has lower
energy, resulting in the hybridized sp orbitals.

from the Na atom to the Cl atom and these atoms gain a plus or minus charge. Both the Na+ and Cl - atoms form
closed shell configurations to form ionic crystals. In III–V compounds like GaAs, both the group-III and group-V
elements have four valence electrons. Then, the bonding configuration is just the same as Si and Ge crystals which
form a diamond structure. However, the bonding configurations contain some ionic features in contrast to the group-IV
elements.

The sp 3 hybrid molecular orbitals and its broadening into bands of tetrahedral and group-VI element semiconductors
are shown in Figure 4.2 10 . When the silicon atoms combine to form a molecule, the hybridized sp 3 orbitals are split
into bonding and antibonding states. These molecular states are broadened into bands due to the electron interaction in
a solid. Thus, in tetrahedral semiconductors, the bonding band forms the valence band and the antibonding band forms
the conduction band. In tetrahedral semiconductors like Si and Ge, the sp 3 hybridization forms the lowest energy
configuration. On the other hand, in group-VI semiconductors like Se, the sp 3 configuration is no longer the lowest-
energy state. The s states lie well below the p states. Only two of the three p states can be utilized for bonding as
shown in Figure 4.2; the coordination number of Se is two. One of the p states is filled with two electrons with
different spins. This is the non-bonding electron pair known as the lone pair. In solid Se, these unshared lone-pair
electrons form a band near the original p-state energy. The bonding and anti-bonding states are split into bands
symmetrically with respect to this lone-pair energy. Therefore, the top of the valence band is formed from the lone-
pair electron states and the bonding p states are placed deeper in the valence band.

In amorphous semiconductors, dangling bonds are formed. As a consequence, a filled state is extracted from the
bonding band and an unoccupied state from the antibonding band. These produce localized states in the gap of
tetrahedral amorphous materials. The structure of amorphous silicon has been analyzed by many methods such as x-
ray diffraction, EXAFS (extended x-ray absorption fine structure), Raman scattering, IR absorption, ESR (electron spin
resonance), NMR (nuclear magnetic resonance), and ENDOR (electron nuclear double resonance).

One of the most important methods for analyzing amorphous semiconductors is the diffraction pattern analysis of x-
ray and its Fourier transform, i.e., radial distribution function (RDF) analysis. The RDF J(R) is related to the pair
distribution function g(R) as follows,
Figure 4.2 Bonding configuration of two amorphous semiconductors: silicon and selenium. From left to right, atomic
states, molecular states, and the broadening of states into bands in the solid are shown.

(4.1)

where p 0 is the average atomic density. The pair distribution function g(R) corresponds to a probability of finding
another atom at distance R from the original atom. In an ideal crystal, g(R) is described by a sum of discrete lines
corresponding to atomic spacing.

The diffraction pattern of an amorphous material is a diffused circular pattern, of which the radial intensity is I(k) for
the scattering wave vector k=4 sin0/ where 0 is the scattering angle and is the wavelength. The Fourier
transform of this intensity function gives the reduced RDF G(r)

(4.2)

Hydrogenated amorphous silicon is characterized by the inclusion of hydrogen in its atomic network. The structural
analysis of a-Si:H by x-ray diffraction patterns revealed the reduced RDF of Equation (4.2) as shown in Figure 4.3 11 .
The general shape of this distribution function shows sharp peaks at small interatomic distances, broader peaks at
larger distances, and then it becomes featureless. The first peak of the RDF (2.36 Å) which corresponds to the Si–Si
bond length is almost the same as that of crystalline silicon, i.e., 2.35 Å. The second peak at 3.86 Å is somewhat
different from the crystalline counterpart of 3.84 Å, but the estimated bond angle of 109° is equal to that of crystalline
silicon. The second and third peaks are apparently much broader in their widths than the first peak because of the
perturbation of the interatomic

Figure 4.3 Radial distribution function of a-Si:H 11 . Short-range order is displayed in the nearest-neighbor atom
location and it is almost the same as that of crystalline silicon, while long-range order is almost completely lost.

distance and bond-angle distribution. In crystalline silicon, the third peak appears at 4.65 Å which is absent in
amorphous silicon. RDF analysis of amorphous semiconductors showed that short-range order (SRO) is maintained,
but long-range order (LRO) is almost completely lacking. From these observations, a u-crystalline model was proposed
to describe the amorphous structure. The u-crystalline structure is a congregation of small-volume crystals. This makes
it easier to understand why SRO is maintained while LRO becomes almost nonexistent. However, this model could not
properly describe the RDF of amorphous semiconductors, so it was abandoned. However, there is a phase of u-
crystalline silicon, if certain fabrication conditions are met. High-temperature or high-power deposition results in this
phase. This is useful for making a film that has low resistivity.

To overcome the difficulty encountered in the u-crystalline description of an amorphous network, many models have
been proposed: an amorphous cluster model, a perturbed crystal model, and a continuous random network (CRN)
model. The CRN model was proposed to explain the apparent contradiction between SRO and LRO. The parameters of
an amorphous material, e.g., interatomic distance and bond angle, are relaxed. These perturbations cause blurring of
the SRO and the disappearance of the LRO. In this model, each atom is bonded to neighboring atoms with their
bonding coordination: four-fold coordination in amorphous silicon. The network is continuous, but it is an "ideal"
amorphous network.

This CRN model was first proposed by Grigorovici, then extended by Polk 12 . Polk's method is based on the following
considerations.

1. The five-member and six-member rings are arranged randomly as starting cores.

2. Dangling bonds are only permitted at the surface.

3. Perturbations in the bond lengths are less than 1% and perturbations in the bond angles are less than 10%.

4. Bond formation is completed to minimize the stress.

In this way, 440 atoms were connected to form a CRN. The resultant RDF showed reasonable agreement with the
experimental data. The lack of LRO is the most prominent feature of this model. It is important to note that the CRN
model corresponds to the ideal amorphous material and does not have defects or dangling bonds. Real amorphous
materials naturally have defects, i.e., vacancies, or interstitial or substitutional impurities. They also tend to have many
dangling bonds since the amorphous solid is deposited under non-equilibrium conditions. Therefore, the CRN model's
applicability in describing the amorphous network is limited, although the model played a significant role in explaining
vitreous semiconductors.

4.2. Density of States of a-Si:H


Generally speaking, the structure of a-Si:H can be understood in terms of a random network model. This network is,
however, disturbed by broken Si-Si bonds, and one hydrogen atom is bound to each of the two Si atoms of the broken
bond. This hydrogen atom compensates for the dangling bond formation and the density of gap states is much lower
than that of conventional amorphous semiconductors. Figure 4.4 shows the schematic representation of the density of
states of a-Si:H. Both the conduction band and the valence band are somewhat diffused compared to the well-defined
band edges shown in Figure 4.2, and form band tails and deep states around the midgap. These tail states correspond
to the irregularities of the atomic configuration. As will be described later in this section, the density of states has a
finite value even in the mid-gap of the energy band. The minimum value is about 1016 cm-3 eV-1 or less. These tail
and mid-gap states, however, are localized states and do not contribute to electrical conduction. In spite of the finite
values of the density of states at Fermi energy E F , many experiments showed that the electrical conduction of
amorphous semiconductors is temperature dependent with an activation energy of E a . To explain
this kind of contradiction, the concept of the mobility edge was introduced. The density of states distribution is divided
into the localized states and the extended states. In the localized states, the electron wave functions are confined to a
small space rather than extended. The greater disorder in the amorphous material disturbs the potential and causes
strong electron scattering. The mobility of electrons in the localized states is therefore assumed to be very low and
only electrons above the mobility edge contribute to conduction. The Fermi energy of doped a-Si:H lies in the band
tails close to the mobility edge, but it never moves into the extended states. No amorphous semiconductors
Figure 4.4 Schematic representation of the density of states and the electron distribution of n-type a-Si:H. The
mobility gap of a-Si:H is 1.9 eV and the optically defined band gap is about 1.7 eV.

have been made with Fermi energy in the extended states above the mobility edge. The closest Fermi energy is about
0.1 eV below the mobility edge. This is the essence of the mobility edge concept and this model explains the
experimental results quite well.

According to this model, the mobility jumps from zero to a finite value at some energy E c, as shown in Figure 4.5.
Above the mobility edge, electron conduction becomes metallic, and this minimum is called the minimum metallic
conductivity (6 min ). Although it is still argued whether this model is valid or not, the concept of mobility edge plays
an important role in understanding the behavior of amorphous materials.

The density of state in the gap of the a-Si:H was first investigated by Spear and LeComber 7 and their field effect
experiments showed that

Figure 4.5 Diagram showing the concept of the mobility edge of the amorphous semiconductors. At the critical energy
of Ec, the mobility increases from zero to a finite value.

the density of localized states is much lower than that of other amorphous semiconductors including evaporated
amorphous silicon. Lang and others applied the technique of deep-level transient spectroscopy (DLTS) to amorphous
semiconductors and studied the density of states quite extensively 13 . The DLTS technology is free from the interface
effect usually encountered in field-effect measurements and the density of states in bulk a-Si:H can be obtained.
However, since the samples used for their measurements are Schottkey diodes, the density of states for the intrinsic or
undoped a-Si:H films is not as easy to evaluate. When the films are doped, however, the density of states can be
measured over a range covering a large part of the energy between the conduction and valence bands.

DLTS was originally introduced to study deep levels in crystalline semiconductors. In principle, it is the synchronous
detection of repetitive and thermally activated transients over a given time constant. The capacitance variation as a
function of time is evaluated at two gate times denoted by t 1 and t 2 . This difference in capacitance plotted as a
function of temperature defines the DLTS spectrum. The DLTS spectrum for
Figure 4.6 Density of states in the a-Si:H gap. Samples are doped with phosphine of which the concentration in
volume ppm in silane is 300 vppm for samples #1 and #3 (10 mol % silane in Ar) and 60 vppm for sample #2 (45
mol% silane in Ar) 13 . The bulk Fermi level of each film is shown by an arrow. The energy levels are normalized to
the conduction-band edge Ec.

amorphous semiconductors can be defined as an extension of this crystalline counterpart, and is considered to be
simply an unresolved superposition of sharp DLTS lines. Figure 4.6 shows the density-of-states for three a-Si:H films
obtained using DLTS technology. The density of states near the mid-gap is observed to vary between values as low as
2×10 15 cm-3 eV-1 in undoped films and as high as 1×10 18 cm-3 eV-1 in phosphorus-doped samples. The general
shape of the density-of-states is dominated by a deep minimum (<10 16 cm-3 eV-1 ) between 0.3 and 0.6 eV from the
conduction band and a broad shoulder of states extending from the valence band to midgap.

4.3. Conductivity and Carrier Mobility


The classical model to describe the electronic transport of amorphous semiconductors relies on the behavior of carriers
near the mobility gap which separates localized and non-localized states. As was discussed in the description of the
density of states in the previous section, the distinction between the localized and extended states is a fundamental
concept of amorphous semiconductors. The carrier transports in localized states are quite limited where a hopping
conduction mechanism prevails. Hopping conduction takes place both in deep levels and in the band tails. The
mechanism is thermally activated and its activation energy equals the hopping energy. This mechanism is important in
amorphous semiconductors that have high density-of-states in the energy gap. In a-Si:H where the density of states is
considerably lower than other amorphous semiconductors, the hopping mechanism is less important and will not be
described here in detail.

However, the concept of the mobility gap is still useful for describing the electronic transport of a-Si:H. The abrupt
increase in the mobility at the mobility edge has not been confirmed at all, and the sharpness of the transition or even
the existence of the edge have been subject to negative opinions. However, even if it is still a hypothesis, it is a good
approximation and can be used as a convenient tool to evaluate the transport behavior of a-Si:H.
The conductivity of a-Si:H is determined from transport in the extended states. In n-type or undoped a-Si:H,
conduction due to holes is generally ignored because of its low mobility so the main contribution to conductivity
comes from electronic transport. The conductivity 6 is given as the product of the carrier density n and the carrier
mobility u,

(4.3)

The summation over the density of states, N(E), results in the following expression, denoting the Fermi distribution
function by f(E):

(4.4)

where the integral is carried out on an electron transport above the Fermi energy, E F . Since the density of state of a-
Si:H is reasonably low in the middle of the gap, and the hopping conduction does not contribute significantly at the
Fermi energy, the dominant conduction mechanism is the electron transport at the conduction band edge. Then,
Equation (4.4) is reduced to

(4.5)

The minimum metallic conductivity introduced in Figure 4.5 can be deduced from Equation (4.5). The conductivity is
written as

(4.6)

where 6 min is given by

(4.7)

where u 0 is the free carrier mobility at an energy of E c and N c is the density of states at E c . This expression is
obtained under an assumption that the mobility edge exists, which has not been established experimentally or
theoretically. However, the conductivity of a-Si:H can be written as in Equation (4.6), in a thermally activated form
over a certain temperature range at least 14 ,

(4.8)

where E a is the activation energy and 6 0 is a prefactor. Figure 4.7 shows the measured relation between the activation
energy and 6 0 for various samples of a-Si:H 15 . This kind of relation is known as the Meyer–Neldel law since they
were the first to observe it in polycrystalline semiconductors. Although there is some scatter in the data of Figure 4.7,
the Meyer–Neldel relation is seen to be obeyed and this relation is written as
(4.9)

where 6 m and T m are constants. From Equation (4.8) and (4.9), 6 m is the conductivity at T=T m . The Meyer–Neldel
relation is a statement that the temperature dependence data for different samples all intersect at the same value of
conductivity, 6 m ~0.1 n-1 cm-1 and at the temperature T m ~600 K15. The explanation for this is still undetermined.

Figure 4.7 Relation between the activation energy and the conductivity constant 60 for doped and undoped a-Si:H
samples 15 . The Meyer–Neldel law is approximately obeyed. The activation energy varies from 1 eV in undoped
material to 0.1 eV in n-type doped samples.

The substitutional doping of a-Si:H was first reported by Spear and LeComber 1 in 1975. This report established the
basis for the subsequent development of a-Si:H technology. Figure 4.8 (a) shows the conductivity of a-Si:H as a
function of the doping gas concentration. The right side of this figure shows the effect of phosphine (PH3 ) doping on
room temperature conductivity. As the phosphine gas is introduced into the
Figure 4.8 (a) Room temperature conductivity and (b) the activation energy of a-Si:H vs. the impurity doping 1 . The
amount of doping is given by the ratio of the phosphine or diborane to the silane , where N
represents the number of molecules per unit volume in the gaseous mixture.

deposition chamber, even a minute quantity of 6 vppm (volume parts per million) can cause conductivity to rise
sharply by two orders of magnitude from undoped conductivity. Further increase in the doping ratio causes the
conductivity to increase by seven orders of magnitude to a conductivity of ~10-2 n-1 cm-1 . The results of transport
experiments clearly showed that conduction takes place predominantly in the extended electron states above EC. The
activation energy, E C-E F is calculated to be 0.6 eV for undoped samples and 0.15 eV for samples prepared at a gas
mixture ratio of 2×10 -3 vppm (see Equation (4.8) and Figure 4.8 (b)).

The left side of Figure 4.8 (a) shows the effect of boron doping. It is interesting to note that the sample remains n-type
up to the gas ratio of about 2×10 -5 . As the diborane ratio is increased above this level, the sample
abruptly becomes p-type and the Fermi level moves past the density of states minimum at E C-E F ~0.85 eV in the
center of the gap. At a doping level of 10-2 , the Fermi energy approaches the valence band up to E C-E
F of about 1.35 eV. Transport now takes place predominantly in the extended hole states at the valence band.

The carrier (electron) mobility is one of the most important parameters in designing an a-Si:H TFT. It is related to the
conductivity through the density of states by Equation (4.4). There are several ways to measure the mobility of a-Si:H.
The time-of-flight technique, SAW (surface acoustic wave) method, and field effect technique utilizing TFT structures
all give consistent results. In these methods the drift mobility is measured rather than the free carrier mobility in the
extended states. The drift mobility u n is a time average of the mobility in the conduction states and the trap states.
This trap-limited transport reflects the density of states distribution in the tail states of a-Si:H. We first consider the
effect of the measured drift mobility u n of a single level of shallow electron traps, where u n is given by the free
carrier mobility in the extended states reduced by the fraction of time when the electrons are trapped in this shallow
level. The electron transit is controlled by multiple trapping and release. On the assumption that the carrier densities
approach the thermal equilibrium distribution, the drift mobility, u n , is expressed as thermally activated by (Equation
(3.13))

where u 0 is the extended state electron mobility, N c is the density at the conduction band edge, n is the total electron
density, and E a is the activation energy.

The time-of-flight measurement scheme to evaluate the drift mobility of a-Si:H is shown in Figure 4.9 16 . This
method was developed by Spear 17 to measure the mobility of materials such as vitreous Se, and can be applied to a-
Si:H that is not heavily doped. The sample structure is quite simple and an a-Si:H thin film with electrodes on both
sides is prepared. The carriers are generated in the neighborhood of one of the electrodes by a short pulse of highly
absorbed light. The voltage is applied to the film such that the holes are collected by the near electrode and the
electrons cross the sample to the opposite electrode and vice versa. As the generated carriers drift a distance x across
the film, they induce a charge Q on the electrode

(4.10)

where t is the sample thickness. The current amplitude is then given by the time derivative of the charge and is
expressed as

(4.11)

Figure 4.9 Schematic illustration of the time-of-flight apparatus used for drift mobility measurements on a highly
resistive specimen.
where v is the drift velocity given by

(4.12)

The current persists until the carriers cross the sample and the transit time is written as

(4.13)

The drift mobility is calculated from Equation (4.13). The results of carrier mobility measurements are shown in Figure
4.10 18 which show the thermally activated behavior of a-Si:H drift mobility.

The drift mobility is also estimated from the current-voltage characteristics of a TFT. The I-V characteristics in the
saturation region (V d >V g –V t) are written as follows (Equation (3.8))

where I d is the drain current, W and L are the channel width and length of a TFT, respectively, C i is the capacitance
per unit area of the gate insulator, and V d , V g , and V t are the drain voltage, gate voltage, and threshold voltage of the
TFT, respectively. This is the same formula as the crystalline silicon MOSFET. Since the structure of the TFT is not
the same as MOSFET (the carriers cross the thin film before going into and getting out of the channel) there is no
proof as to whether this formula can be applied to the TFT. However, it is well established that the TFT behavior,
especially a-Si:H TFT behavior, is reasonably well described by this model. Therefore, Equation (3.8) can be used to
estimate the drift mobility of a-Si:H (from a linear plot of vs. V g ). It is worth noting that the drift mobility
obtained from the TFT characteristics reflects the interface property between the a-Si:H and the gate insulator. Great
care has to be taken in preparing the gate insulator and selecting the insulator material.

The surface acoustic wave (SAW) method 19 is a useful technique for estimating the drift mobility of doped samples,
in contrast to other techniques, such as the time-of-flight technique, where the samples are limited to lightly doped or
undoped films. In the SAW method, the sample is placed a short distance, h, above a piezoelectric crystal like LiNbO 3
as shown in Figure 4.11. A surface acoustic wave is propagated in the LiNbO 3 crystal which causes a fringe field in
the a-Si:H film. This field produces a voltage between the two electrodes separated in
Figure 4.10 (a) Electron and (b) hole drift mobility data points from the time-of-flight experiments 18 . The lines are
the theoretical fits to the multiple trapping transport mechanism with an assumed exponential band tail. The carrier
mobilities at the conduction band edge and at the valence band edge are, respectively, 13 cm2 /Vs (electrons) and 0.5
cm2 /Vs (holes).

Figure 4.11 Cross-sectional view of the travelling wave experiments 19 . The a-Si:H film is located above the LiNbO 3
plate at a distance of about 1 um or less. The gap is filled with dry He gas.

the direction of the travelling wave. The acousto-electric voltage V ae is given by

(4.14)

where L is the electrode spacing, and v s , k a , and 0 are the velocity, wave number, and electric potential amplitude
of the traveling wave at the surface of the LiNbO 3 plate. The factor g is called the attenuation coefficient and relates
the fields inside the semiconductor to the potential at the surface of the piezoelectric crystal. This factor is given by

(4.15)
where t is the a-Si:H film thickness and A and B are coefficients of the complex electric potential in the
semiconductor and are functions of the conductivity of a-Si:H, the SAW frequency, the gap (h), and the dielectric
constants of the semiconductor and glass substrate.

The mobility value measured by the SAW method agreed well with the calculated values from the multiple trapping
model of Equation (3.13). For this measurement, only those carriers trapped in the transport states at least once during
the SAW period can participate in the charge bunching in the traveling wave. In this sense, their thermal release rate v
0 exp[-(E c-E/kT] must be equal to, or greater than the traveling wave frequency v. Therefore, the energies of these
trapped carriers must lie between the conduction band edge E c and the demarcation energy E d given by

(4.16)

If the electron escape frequency is assumed to be v 0 =5×10 12s -1 , E c -E d has values between 0.24 and 0.42 eV at
temperatures between 300 and 430K for frequencies in the 10 to 100 MHz range.

All these measurements by the time-of-flight, the field-effect, and the SAW methods give consistent electron mobility
values which are in the range of 0.3 to 1 cm2 /Vs at room temperature. The free electron mobility u 0 of a -Si:H is
estimated to be in the range of 5 to 20 cm2 /Vs. The free hole mobility is estimated as about 1 cm 2 /Vs.

4.4. Absorption Spectra of a-Si:H


4.4.1. Optical Absorption

Optical absorption comes from electronic transitions between the valence and conduction bands of semiconductors.
The absorption edge corresponds to the minimum energy difference E g between the lowest minimum of the
conduction band and the highest maximum of the valence band. The k (momentum) conservation rule has to be kept
during transitions in crystalline semiconductors. If a transition occurs with the same k value, direct transition if
possible. But, for transitions between different k values, the emission or absorption of a phonon compensates for the
difference in the momentums (indirect transition).

In amorphous semiconductors, however, there is no distinction between a direct and indirect gap, since the momentum
is not conserved. The absorption constant a is calculated by multiplying the matrix element and the joint density of
state of the valence and conduction bands. According to Tauc 21 , the absorption edges of amorphous semiconductors
are divided into three regions as shown in Figure 4.12: the high absorption region, the exponential part, and the weak
absorption tail in the low energy region. This type of behavior is quite common in many amorphous semiconductors.

Assuming that the momentum conservation rule is relaxed and that the matrix elements for the electronic transitions are
constant over the
Figure 4.12 Three regions of the absorption edge of amorphous semiconductors 21 . Region A corresponds to the Tauc
edge. The extrapolation of a 1/2vs. energy in this region gives the optical gap of the amorphous material. Region B
represents the exponential tail called the Urbach edge. The region C is the weak absorption tail.

photon energies of interest, we obtain the absorption edge formulation as follows. 22 We consider only the transitions
between a localized state and an extended state, and ignore the transitions between localized states. The absorption
constant at a frequency of , a( ), is expressed as a function of conductivity 6( )

(4.17)

when n is the refractive index and c is the speed of light.

The conductivity 6( ) at a frequency of is given by

(4.18)

where V is the volume, p is the matrix element, m is the electronic mass, h is Planck's constant, and N c (E), and N v
(E) are the densities of state of the conduction and valence bands, respectively.

The matrix element p for transitions between states in different bands is assumed to be the same as that for transitions
between extended states in the same band. In this case, the matrix element is given by

(4.19)

where a is the average lattice spacing, and Equation (4.17) reduces

(4.20)

where integration is over all pairs of states in the conduction and valence bands separated by energy h .

If we assume that the density of states at the bottom of the conduction band is expressed by N c (E) (E-E A ) r 1 and
at the bottom of the valence band by N v (E) (E B - E) r 2 , where E A and E B are as shown in Figure 4.4, then it
follows from Equation (4.20) after some manipulation 22

(4.21)

where =E A -E B and C is a constant. Under the assumption of parabolic bands, i.e., r 1 =r 2 =1/2, which still needs
justification, Equation (4.21) reduces to

(4.22)

The absorption edge region A of Figure 4.12 corresponds to the absorption coefficient a( ) of Equation (4.22) and is
known as the Tauc edge. Figure 4.13 shows an example of (ah ) 1/2 of a-Si:H plotted as a function of energy 23 .
The relation expressed in Equation (4.22) has been
Figure 4.13 Tauc plot of a-Si:H. The extrapolated (optical) band gap is 1.7 eV 23 .

observed to apply to the absorption of many amorphous materials. The extrapolation of this plot gives a well-defined
optical band gap of 1.7 eV for a-Si:H. The mobility gap, on the other hand, is defined as the energy separation of the
conduction and valence band mobility edges. The mobility gap is generally larger than the optical gap and is 1.9 eV
for a-Si:H.

Below the Tauc edge, there is an exponential region of the absorption edge. In this region, the absorption coefficient is
described by the formula

(4.23)

where E G is the energy to the order of and E O is the characteristic energy representing the logarithmic slope of
the absorption coefficient in this regime. The exponential tail of Equation (4.23) is called the Urbach edge after the
first person to make the observation in alkali halides. The energy E O in amorphous semiconductors is constant in the
temperature range from 300K up to near the glass transition temperature T g , although Eo in alkali halides is
temperature dependent near 300K and is given by about kT/0.8.

The Urbach edge of a-Si:H is shown in Figure 4.14 24 and an exponential tail is observed for an absorption coefficient
range between 1 and 1,000 cm-1 . Since the Urbach edge is observed in both crystalline and amorphous
semiconductors, there seems to be no simple explanation how this exponential tail occurs. The fundamental basis for
this phenomenon is still controversial.

4.4.2. IR Absorption
The a-Si:H films prepared in a glow-discharge apparatus contain bonded hydrogen of about 10 atomic percentages in
the silicon atom matrix. This incorporation of hydrogen helped reduce defects due to silicon dangling bonds. The
presence of hydrogen which makes electronic-grade a-Si:H possible can be detected by using techniques such as
hydrogen evolution, IR absorption, and Raman spectroscopy. The evolution experiment reveals the quantitative data of
hydrogen, but it does not give us any information on how hydrogen is incorporated into the silicon network.

Information on the bonding configuration between silicon and hydrogen is obtained by infrared absorption
spectroscopy pioneered by Brodsky et al 25 . The local bonding arrangements of hydrogen atoms in a-Si:H alloys are
classified into SiH, SiH2 , SiH3 , and polyhydride (SiH 2 )n 25 , 26 . Of these arrangements, the SiH configuration is the
most important as device grade a-Si:H, and the nomenclature of a-Si:H first coined by Brodsky et al. refers to an
amorphous network only where hydrogen atoms are bonded to separate silicon atoms or the structural arrangement is
predominantly SiH. In other cases, where SiH2 and SiH3 grouping or even polymerization are favored, they refer to the
network as amorphous silicon hydride or a-SiH x .

The SiH environment of an amorphous network can be clearly confirmed by measuring the IR absorption and Raman
spectroscopy. For IR transmission measurements, a-Si:H films are deposited on high-

Figure 4.14 Absorption edge of a-Si:H in the Urbach region 24 . The film with the higher Tauc gap was grown at a
substrate temperature of 250°C compared with 200°C for the lower gap film. The hydrogen contents are 14 at. % and
9 at. % for the higher and lower gap film, respectively.

resistivity (>100 n cm) single-crystal silicon wafers polished on both sides. The index of refraction of crystalline
silicon matches the index of refraction of a-Si:H better than most other IR transmitting materials. Therefore,
interference fringes from multiple reflections within the film are minimized. The silicon wafers are wedge-shaped for
the same purpose.

Figure 4.15 shows the IR transmission of a-Si:H samples deposited by glow-discharge dissolution at a substrate
temperature of 230°C 26 . These spectra depend on the deposition conditions. However, the three peaks shown in the
figure almost always show up. These peaks correspond to the phonon frequencies based on silicon-hydrogen bonds. To
convert the transmittances to absorption coefficients, the interference-free transmittance T is approximated by

(4.24)

where a is the absorption coefficient, t is the film thickness, and R is an empirically determined interface multiple
reflection loss. The value of R is determined by setting T=T 0 when a=0. For the transmission of the silicon substrate
used, T 0 =0.54. Then, Equation (4.24) reduces to

(4.25)

Figure 4.15 IR transmission of a-Si:H for the samples prepared at a substrate (anode) temperature of 230°C and under
a fixed dilution of SiH4 by Ar: 5 at. % 26 .

where it was assumed that the refractive indices of a-Si:H (n of about 3.0) are equal to those of the silicon substrate (n
of about 3.4), while keeping accuracy within ±10%.

Once the dependence of the absorption coefficients on frequency is known, the concentration of hydrogen N H can
be calculated from
(4.26)

where

(4.27)

where s / o is the relative dielectric constant (for a-Si:H; s / o =12) and C is a constant which depends on the Si-
Hn vibrational modes and the hydrogen grouping 25 . For example, the constant is C=1.72×10 20 cm-2 for the stretching
absorption band of Si-H bonds. Hydrogen concentration of between 10 and 15% has been estimated from the analysis
which is consistent with estimates made using other technologies such as hydrogen evolution. Three groups of
absorption bands are in the IR spectra as shown in Figure 4.15. There are two strong bands: one at 2000 to 2100 cm-1
and the other at 630 to 640 cm-1 . There is a weaker peak at the 850 to 900 cm -1 band. The assignment of these peaks
and bands to a specific vibrational mode was carried out by analyzing the IR spectroscopy and Raman spectroscopy as
well as their dependence on parameters such as deposition temperature, RF power, and hydrogen concentration. The
results are summarized in Figure 4.16 25 , 26 . The strong absorption bands at 2000 to 2150 cm-1 correspond to the
stretching modes. In the case of monohydride (SiH), the absorption occurs at 2000 cm-1 while the bands shift to higher
frequencies of 2090 to 2150 cm-1 for SiH2 and SiH3 . The other strong bands are for the rocking and wagging modes in
the frequency range of 630 to 650 cm-1 . They form rather broad bands and the differences among SiH, SiH2 , and SiH3
groupings are not recognized explicitly. The bands in the 850 to 910 cm -1 range correspond to the bending (twist and
deformation) modes.

References
1. Spear, W.E. and LeComber, P.G. (1975). Substitutional doping of amorphous silicon. Solid State Communications ,
17, 1193–1196.
Figure 4.16 Schematic illustration of the vibrational modes of SiH, SiH2 , and SiH3 groupings in a-Si:H 25 , 26 . The
solid circles represent Si atoms and the open circles represent H atoms. The wavenumber shown for each mode
corresponds to the IR wavenumber in cm-1 .

2. Carlson, D.E. and Wronski, C.R. (1976). Amorphous silicon solar cell. Applied Physics Letters , 28, 671–673.

3. Tawada, Y., Tsuge, K., Kondo, M., Okamoto, H., and Hamakawa, Y. (1982). Properties and structure of a-SiC:H
for high-efficiency a-Si solar cell. Journal of Applied Physics , 53, 5273–5281.

4. Yamamoto, H., Baji, T., Matsumaru, H., Tanaka, Y., Seki, K., Tanaka, T., Sasano, A., and Tsukada, T. (1983).
High-speed contact type linear sensor array using a-Si pin diodes. In Extended Abstracts of the International
Conference on Solid State Devices and Materials (Tokyo, 1983) pp. 205–208.

5. LeComber, P.G., Spear, W.E., and Gaith, A. (1979). Amorphous-silicon field-effect device and possible application.
Electronics Letters , 15, 179–181.

6. Snell, A.J., Mackenzie, K.D., Spear, W.E., and LeComber, P.G. (1981). Application of amorphous silicon field
effect transistors in addressable liquid-crystal display panels. Applied Physics , 24, 357–362.

7. Spear, W.E., and LeComber, P.G. (1972). Investigation of the localized state distribution in amorphous Si films.
Journal of Non-Crystalline Solids , 8–10, 727–738.
8. Chittick, R.C., Alexander, J.H., and Sterling, H.F. (169). The preparation and properties of amorphous silicon.
Journal of Electrochemical Society , 116, 77–81.

9. Mott, N.F. (1969). Conduction in non-crystalline materials III. Localized states in a pseudogap and near extremities
of conduction and valence bands. Philosophical Magazine , 19, 835–852.

10. Kastner, M. (1972). Bonding bands, lone-pair band, and impurity states in chalcogenide semiconductors. Physical
Review Letters , 28, 355–357.

11. Schülke, W. (1981). Structural investigation of hydrogenated amorphous silicon by X-ray diffraction.
Philosophical Magazine B , 43, 451–468.

12. Polk, D.E. (1971). Journal of Non-Cystalline Solids , 5, 365.

13. Lang, D.V., Cohen, J.D., and Harbison, J.P. (1982). Measurement of the density of gap states in hydrogenated
amorphous silicon by space charge spectroscopy, Physical Review B , 25, 5285–5320.

14. Street, R.A. (1991). Hydrogenated Amorphous Silicon , pp. 227–229. Cambridge: Cambridge University Press.

15. Tanielian, M. (1982). Adsorbate effects on the electrical conductance of a-Si:H, Philosophical Magazine , B45,
435–462.

16. Street, R.A. (1991). Hydrogenated Amorphous Silicon , pp. 75, Cambridge: Cambridge University Press.

17. Spear, W.E. (1969). Drift mobility techniques for the study of electrical transport properties in insulating solids.
Journal of Non-Crystalline Solids , 1, 197–214.

18. Tiedje, T. (1984). Information about band-tail states from time-of-flight experiments. In Semiconductors and
Semimetals , edited by R.K. Willardson and A.C. Beer, Vol. 21: Hydrogenated Amorphous Silicon, Part C, Chap. 6.
Orlando: Academic Press.

19. Fritzsche, H. (1984). Analysis of the traveling-wave technique for measuring mobilities in low-conductivity
semiconductors, Physical Review B , 29, 6672–6678.

20. Kaneko, Y. and Fritzsche, H. (1991). Frequency dependence of drift mobility in a-Si:H measured by traveling-
wave method, Journal of Applied Physics , 69, 8237–8240.

21. Tauc, J. (1974). Optical properties of amorphous semiconductors. In Amorphous and Liquid Semiconductors ,
edited by J. Tauc, P. 172. London and New York: Prenum Press.

22. Mott, N.F. and Davis E.A. (1979). Electronic Processes in Non-Crystalline Materials , 2nd edn, pp. 287–289.
Oxford: Clarendon Press.

23. Cody, G.D. (1984). In Semiconductors and Semimetals , edited by R.K. Willardson and A.C. Beer, Vol. 21, Part B,
Chapter 2. Orlando: Academic Press.

24. Cody, G.D. (1992). Urbach edge of crystalline and amorphous silicon: a personal view. Journal of Non-Crystalline
Solids , 141, 3–15.

25. Brodsky, M.H., Cardona, M., and Cuomo, J.J. (1977). Infrared and Raman spectra of the silicon- hydrogen bonds
in amorphous silicon prepared by glow discharge and sputtering. Physical Review B . 16. 3556–3571.

26. Lucovsky, G., Nemanich, R.J., and Knights, J.C. (1979). Structural interpretation of the vibrational spectra of a-
Si:H alloys. Physical Review B , 19, 2064–2073.
CHAPTER 5
Twisted-Nematic Cell
The electro-optical effect in a twisted nematic (TN) liquid-crystal cell described in 1971 1 is now widely used in
active-matrix liquid-crystal displays. As shown in Figure 5.1, a thin (5 um) layer of nematic liquid-crystal is placed
between two glass plates provided with a transparent and conductive coating. On both plates the orientation of liquid-
crystal molecules or director, is aligned to be nearly parallel to the surface of the glass plate (planer or homogeneous
alignment), and as shown schematically in Figure 5.2 (a), on each plate this orientation is twisted 90° with respect to
that of the other plate, so that the orientation is continuously twisted from the bottom to the top glass plate by 90°. The
homogeneous alignment of liquid-crystal molecules on each plate is produced by rubbing the surface of a polyimide
thin film in the proper direction with

Figure 5.1 Fundamental configuration of the TN cell. A thin ( 5 um) layer of liquid crystal is placed between two
glass plates. The cell is driven by an ac voltage applied to the transparent electrodes of ITO. The planar or
homogeneous molecular alignment is achieved by the rubbed surface of the polyimide film, and the alignment
orientations on the two glass plates are perpendicular to each other.
Figure 5.2 (a) Molecular alignment in the TN cell. When no electric field is applied, the orientation of molecules
rotate 90° from the bottom glass plate to the top glass plate. Linearly polarized light is also rotated by 90° as it travels
through the cell.

a fabric. This TN orientation pattern is analogous to the planar texture of cholesteric liquid-crystals (Chapter 6). With
a twist angle of 90° and a cell thickness of 5 um, the resulting pitch of a cholesteric equivalence is 20 um. Since this is
much greater than the wavelength of light, the polarization plane of linearly polarized light travelling normal to the
Figure 5.2 (b) When a sufficiently high field is applied to the cell, the molecules are aligned perpendicular to the glass
(homeotropic alignment) and the rotation of the polarized light does not take place. Light is thus blocked by the top
polarizer and does not come out of the module.

glass plates rotates with the liquid-crystal axis. The 90° twist of the director should then lead to a 90° rotation of the
linearly polarized light. Therefore, in this normally white (NW) mode of operation where the two polarizers are set
perpendicular to each other, the light is transmitted through the cell when there is no applied voltage.

A voltage applied to the cell modifies the director orientation. When the applied voltage is below a threshold level
there is no change of the orientation, but at the threshold the molecular orientation begins to be aligned to the electric
field and tends to be perpendicular to the glass plate. At voltages well above the threshold, the alignment of the
molecules is completely parallel to the field except the regions adjacent to the surface of the glass (homeotropic
alignment). In this state, shown in Figure 5.2 (b), the TN cell becomes optically inactive and linearly polarized light
travels through the cell without any rotation of the polarization. Then, the light polarization becomes perpendicular to
the output polarizer resulting in no transmission of light.

The originally proposed TN cell used a parallel scheme: the two polarizers were parallel to each other. In this NB
(normally black) configuration, the light is not transmitted in a field-free state and is transmitted when the applied
voltage is above the threshold. When there is no bias voltage, however, transmission is suppressed to zero only for
monochromatic light of a wavelength of 2d Therefore, in a practical display in which a broad-band
illumination is used, there is a small amount of light leakage as shown in Figure 5.3. In the NB mode this leakage
reduces a contrast ratio, which is the ratio between the "on" and "off" transmissions.
Figure 5.3 Normally black (NB) and normally white (NW) modes of operation. These modes respectively correspond
to the parallel and the perpendicular polarizer schemes.

In the normally white mode of operation, however, there is little reduction in the contrast ratio since this ratio is
governed by an "off" transmission corresponding to a high bias voltage. When a sufficiently high voltage is applied,
the liquid-crystal molecules are aligned parallel to the electric field or perpendicular to the glass substrate, and there is
no rotation of electric vector of the polarized light. Transmission is therefore suppressed completely regardless of the
wavelength of the light. The transmission in the "on" state is wavelength dependent, but this transmission does not
have a critical effect on the contrast ratio. A contrast ratio exceeding 100 can be readily obtained in the normally white
mode of operation. This is why the NW mode is preferred in the practical TFT/LCD display.

The threshold voltage (V th) of the TN cell is defined as the extrapolated voltage of the transition region from the
homeotropic to the homogeneous alignment as shown in Figure 5.3. The V th of the cell is designed to be 2–3 V, and is
determined by the choice of the liquid-crystal materials to be mixed. The maximum operating voltage, or the peak
amplitude of the data signal of the display, is about 7–8 V. This low voltage and the low current due to the high
resistance of the liquid-crystal contribute very much to the low power consumption of the display. The current density
of the working display is less than 0.1 uA/cm2 and the power level is therefore on the order of 1 uW/cm 2. The power
required for a 10" display is less than 300 uW as far as the liquid-crystal is concerned.

5.1. Threshold Voltage of TN Cell


The liquid-crystal molecules are easily affected by an external field. In a cell structure like that of a TN cell, the
molecules at the boundaries are strongly or loosely coupled to the surfaces of parallel plates. Before considering the
threshold voltage of a TN cell, let us consider a simple case 2 (Figure 5.4) in which a nematic layer of thickness d is
between the planes (z=0 and z=d) of a Cartesian coordinate system. The directors at the boundaries are not influenced
by the external fields, always lie in the xy plane, and are assumed to be parallel to the x axis. With the external electric
field E applied, the tilt angle between the director and the xy plane is expressed by 0. This modification of 0(z) is
induced
Figure 5.4 Simple configuration of a nematic cell in which the director orientation is confined to the x-z plane and is
parallel to the x-axis at the boundaries.

when the external field exceeds some critical value. This phenomenon is called the Freedericksz transition 3 and is
observed in such cell arrangements as homogeneous alignments, homeotropic alignments, and their hybrids.

In a homogeneous alignment with the director parallel to the x axis, the director with the applied external field is
expressed as

(5.1)

whereas the director without a field is expressed as

(5.2)

The boundary conditions of strong anchoring on the surfaces of glass plates are give by

(5.3)

where d is the cell spacing.

In this configuration the free energy F per unit area of curvature strain is given by

(5.4)

where /_\ = // - - is the dielectric anisotropy and k 11 and k 33 are the elastic constants of the liquid-crystal (see
Chapter 6). From the condition that this free energy takes a minimum value, we obtain the following equation (Euler's
equation)
(5.5)

For small 0 this equation reduces to

(5.6)

and, for /_\ >0, the solution of this equation is written as

(5.7)

where C 1 and C 2 are constants. The boundary conditions (0(0)=0(d)=0) yield

(5.8)

(5.9)

From Equation (5.9) we obtain

(5.10)

The case of m=0 corresponds to that in which there is no external field. The angle 0 starts to tilt at the threshold field E
th , and this corresponds to the case of m=1. For /_\ >0, E th is given by

(5.11)

and the threshold voltage is given by 4

(5.12)

We can see that the threshold voltage is independent of the cell gap, a very important conclusion.

For the case of finite 0, the integration of Equation (5.5) yields

(5.13)
We seek a solution of the type d0/dz=0 at z=d/2, since 0(z) takes a maximum 0 m at z=d/2 and is symmetric with
respect to this plane. The constant C 3 is then given by

(5.14)

Equation (5.13) is reduced to

(5.15)

Thus,

(5.16)

This equation gives the relation between the maximum tilt angle 0 m and the electric field. Figure 5.5 shows an
example of this relation when k 33/k 11=1 5 .

Figure 5.5 Maximum tilt angle vs. the electric field normalized to the threshold in the simple cell configuration of
Figure 5.45. The maximum tilt angle occurs at the midplane of the cell.
Figure 5.6 Cartesian coordinate system for analysis of the twisted-nematic cell. The director orientation at the bottom
(z=0) is assumed to be parallel to the x-axis ( (0)=0). At the top (z=d), it is twisted from the x-axis by 0 , which is
generally equal to /2 ( (d)= /2).

Next, we consider the threshold voltage of a TN cell. The liquid-crystal in a TN cell is sandwiched between the two
glass plates and the director orientation at the surfaces of these plates differ from each other by an angle of 0 , which
commonly takes a value of /2. The director orientation is a function of z (in the coordinate system shown in Figure
5.6) and is parallel to the x axis at z=0. And the director at z=d is twisted by 0 from the x axis. When there is no
external field the director is written as n =[cos (z), sin (z), 0], and the boundary condition is (0)=0, and (d)=
0 . When an external electric field is applied parallel to the z axis, the director is affected by this field and tends to be
aligned to the field. Writing the deviation from the x-y plane as 0(z), we obtain

(5.17)

The boundary conditions of 0(z) are again those of a strong anchoring: 0(0)=0(d)=0. The free energy F of this system
per unit area is then given by

(5.18)

where k 11 , k 22, and k 33 are respectively the constants of splay, twist, and bend stresses. From the condition that the
free energy takes a minimum value, we obtain the following equations (Euler's equation):

(5.19)

and

(5.20)

For small 0(z), Equations (5.19) and (5.20) are linearized to


(5.21)

and

(5.22)

From Equation (5.21) and the boundary condition, we obtain

(5.23)

and Equation (5.22) is thus reduced to

(5.24)

Following the same procedure as with the simple nematic cell, Equation (5.6), we obtain the threshold voltage V th of
the TN cell as

(5.25)

for 0= /2 and /_\ >0. This threshold voltage is again independent of the cell gap.

The response time of the TN cell is derived from the equation of motion of the director, which describes the balance
between the torques due to the elastic and viscous forces and the external field. Equating the left-hand side of
Equation (5.24) to Y 1 (d0/dt), we obtain

(5.26)

where t is the time and Y 1 is the coefficient of the rotational viscosity (see Equation (6.72)). Equation (5.26) holds
only for small 0 but is adequate for the present purpose. If we assume

(5.27)

where is the rise-time constant, Equation (5.26) is reduced to

(5.28)

where we further assumed for small t. The same procedure as in Equation (5.24) leads us to the next
equation
(5.29)

From Equation (5.29), the time constant is given as

(5.30)

where V th is given by Equation (5.25).

The decay-time is obtained following the same procedure:

(5.31)

We can see from Equations (5.30) and (5.31) that the time constant is proportional to the square of the cell gap.
Reducing the cell gap is therefore an effective way to shorten the cell's response time 6 .

For a display like an LCD as well as a CRT, the rise and fall times are defined for convenience as the times when the
brightness reaches, after the turn-on and turn-off of the drive voltage, 90% and 10%, respectively of full brightness
(Figure 5.7). These turn-on and turn-off times (t ON and t OFF) are given by

(5.32)

and

(5.33)

An example of the temperature dependence of t ON and t OFF is shown in Figure 5.8 7 . The turn-on time is usually
shorter than the turn-off time, and at room temperature these times are of the order of 10 ms. The response time is
reasonable for such application as TVs and computer terminals.

5.2. C-V Characteristics of TN Cell


As described in Section 2.3.2, the dc voltage offset appears on the pixel electrode because of the parasitic capacitance
of the a-Si TFTs. This offset /_\V is given by (Equation (2.30))
Figure 5.7 Response time of a twisted-nematic liquid-crystal cell. At t=t ON after the ac voltage turn on, the display
brightness reaches 90% of its full value. At t= t OFF after the voltage cutoff, the brightness becomes 10% of the full
value.

Figure 5.8 Temperature dependence of the response time of the TN cell with a 4.8-um cell gap 7 . At room
temperature (T=20°C), Ton=14 ms and Toff =17 ms.
Figure 5.9 Dependence of the liquid-crystal cell capacitance and NW cell transmittance on the applied voltage.

This offset voltage will be easier to cancel out if the capacitance of the liquid-crystal cell, C lc is constant.
Unfortunately, C lc depends on the applied voltage and the dc voltage offset changes from pixel to pixel and from time
to time.

The variation of C lc with the bias voltage as shown in Figure 5.9 is due to an anisotropic dielectric constant of liquid-
crystal. When the bias voltage is lower than the threshold voltage, the capacitance of the twisted-nematic liquid-crystal
cell is given by

(5.34)

where d is the cell gap, A is the effective area of the pixel electrode, and - is the dielectric permittivity for the
direction perpendicular to the director. Above threshold, the capacitance begins to increase from the value of Equation
(5.34). When the bias voltage is sufficiently high, C lc is given by

(5.35)

where // is the dielectric permittivity for the direction parallel to the director. It is worth noting that the dielectric
permittivity used here is for the low frequency electric field. Because the electric field is applied perpendicular to the
glass substrate, it is perpendicular to the director or the optical axis of the liquid-crystal below threshold. On the other
hand, the electric vector of the linearly polarized light is parallel to the glass and parallel to the director. Therefore,
below threshold the dielectric permittivity is = // for the optical wavelength, whereas it is = - for the static
field. The situation is reversed, however, well above the threshold. As shown in Figure 5.2 (b), the director of the
liquid-crystal becomes vertically aligned and the dielectric permittivity becomes = - for the optical wavelength and
= // for the static field.

5.3. Optical Properties of TN Cell


The TN cell consists of the polarizer, the liquid-crystal (TN), and the analyzer. The polarizer (analyzer) is a device that
transmits light with its electric field in one direction and absorbs light with its electric field in other directions. The
electric field of light is perpendicular to the direction of propagation (z-axis). The magnetic field of light is
perpendicular to both the electric field and the direction of propagation. The magnitude of the magnetic field, however,
is E/c (where c is the speed of light). Therefore, the characterization of light is mainly based on the electric field. The
electric field oscillates with a certain frequency. The x and y components of the electric field oscillate sinusoidally
with this frequency but with an independent phase. If the phase difference is p, the electric field is expressed as
follows:

(5.36)

and

(5.37)

where =2 f is the angular frequency. Linearly polarized light is a phase-matched light: p=0 (or p=n ).
Circularly polarized light (Figure 5.10), is the light where the phase difference is equal to /2 (or p=(2n-1) /2) and
the amplitudes of the x and y components are the same (E xo =Eyo ).

The molecules in the TN cell undergo a 90° rotation from the top to the bottom of the liquid-crystal cell. The material
used in these devices is birefringent and has a positive dielectric constant; that is,

(5.38)

(5.39)

Figure 5.10 Circularly polarized light has x and y components with the same amplitude and the phase difference is a
multiple of /2.

where n o and n e are the ordinary and extraordinary refractive indices. In a liquid-crystal this birefringence is given by

(5.40)

The 90° rotation of molecules in this system is associated with the rotation of the plane of polarization of light
transmitted through the cell. If polarized light is incident on the cell in such a way that the polarization is parallel to the
director at the entrance, its polarization is rotated 90° as it exists the liquid-crystal (Figure 5.2 (a)).

When the electric field is applied along the z axis, the tilt angles of the molecules are modified to become realigned
parallel to the field. If the field becomes high enough, the molecular realignment is completely parallel to the field and
forms the homeotropic state which has no optical activity. Between crossed polarizers as shown in Figure 5.2, this TN
device exhibits a transmitting behavior (normally white) when there is no applied field (Figure 5.2 (a)) and is switched
to an extinction state when an electric field is applied (Figure 5.2 (b)).

The propagation of the polarized light in an anisotropic medium like a liquid-crystal can be analyzed by solving the
equation that governs the evolution of the ellipse of polarization of a light wave 8 , 9 . The equation is a first-order
ordinary differential equation for = (z, 0 ), where 0 is the value at z=0; that is, the incident state of
polarization. is a single complex variable which describes the ellipse of polarization of the incident light wave. Both
the azimuth and ellipticity e can be calculated from as follows 10 .

(5.41)

(5.42)

In a TN cell where the director is parallel to the x axis at z=0 and it is aligned to the y axis at z=d, the polarization is
described by the following equation.

(5.43)

The general solution of this equation is obtained by putting it in the form

(5.44)

Then the polarization at the exit plane (z=d) is expressed as

(5.45)

where u=2d /_\n/ for the total twist angle of /2.

From Equations (5.41), (5.42), and (5.45), we obtain

(5.46)

(5.47)

These equations define the polarization state of the radiation emerging from the TN cell. Figure 5.11 shows the
ellipticity e of Equation (5.46) and the rotation angle of z(=( /2)+ ) of Equation (5.47) as a function of the reduced
cell gap u (=2/_\ n d/ ). Both the ellipticity and the angle of rotation oscillate with u, but tend to be independent of u
when it is large. In the limit of large u,
(5.48)

the polarization of the incident light is rotated by exactly 90° as the light traverses the cell and, regardless of its
wavelength, is linearly polarized to the y axis when it exits the cell. This condition is called Mauguin's limit. To reach
this limit, however, an impractically large value of d is necessary; and in practice a smaller value of u is chosen. In
this case, the emitted light becomes elliptically polarized as it exits

Figure 5.11 Ellipticity e and the rotation angle z (= ( /2)+ ) at the exit plane of a TN cell. The most common
choice of u (=2/_\nd/ ) for a display cell is u= .

the cell. The exiting light is linearly polarized only when the following equation is satisfied:

(5.49)

where m is an integer. The smallest and most commonly used cell gap is for m=2 and is given by

(5.50)

The transmittance of radiation, T, is obtained from Equation (5.45) and is given for a system with parallel polarizers
(NB cell) as

(5.51)
and for a system with crossed polarizers (NW cell) as

(5.52)

Figure 5.12 Transmission of the TN cell as a function of u=2/_\nd/ . The curves T and 1-T respectively correspond
to the transmittances of the normally black cell and the normally white cell.

Equations (5.51) and (5.52) are plotted in Figure 5.12.

As before, the transmission oscillates with u; but for a u value corresponding to Equation (5.50) there is no
transmission of light for parallel polarizers and 100% transmission of light for crossed polarizers when no field is
applied. For a given cell gap of d, however, this condition is achieved only for the monochromatic light with a
wavelength given by

(5.53)

In TFT/LCDs, three colors (R, G and B) are used and it is not possible for one cell to satisfy Equation (5.53) for all
these colors. And because each color has a bandwidth of about 100 nm, it is not possible to satisfy Equation (5.53)
even for one color. For a wavelength other than that given by Equation (5.53) the emitted light is elliptically polarized
and the optical axis does not coincide with they-axis. Figure 5.13 shows the transmission of the display panel vs. drive
voltage for the NB and NW cells 11 . In the NB or parallel polarizer system, there is a substantial leakage of light when
no field is applied and a scatter in the transmission is observed for the R, G and B colors. Since the contrast ratio CR
of the display is defined as

(5.54)
Figure 5.13 Transmittances of NW and NB TN cells as a function of the applied voltage 11 . A /_\nd value of 0.5 um
is most commonly used in the practical display. The wavelength dependence of the transmission below threshold is
critical in determining the contrast ratio of the NB cell.

where T on and T off are the full transmittance and the off transmission, the leakage and scatter in the off-state
transmission greatly lower the contrast ratio. This is why the crossed-polarizer system or the NW mode operation is
preferred in a practical display.

Even in the NW-mode.operation shown in Figure 5.13, the scatter in the transmission when no field is applied is
obvious. In this operational mode, however, the contrast ratio can be made high by increasing the maximum operating
voltage. When the voltage is increased, the transmission is reduced and its wavelength dependence disappears. The
influence of the scattering of the field-free transmission on the contrast ratio is not large and a contrast ratio higher
than 100 can be obtained relatively easily over a full color range. Because of this effect, the normally white mode is
preferred and is used most widely.

The viewing-angle characteristics of the display are also important. The transmission shows complex behavior as the
viewing angle is changed, since the optical path length depends on the angle of incidence. D.W. Berreman 12
calculated the viewing-angle characteristics of the cell for the coordinate system of Figure 5.6. The tilt angle 0 and the
azimuthal or turn angle are defined there and the director at the bottom surface is oriented parallel to the x-axis and
the director at the top surface lies parallel to the y-axis. Light rays travel along the z-axis. The electric field is applied
in the z-direction.

The tilt and turn angles 0 and are shown in Figure 5.14 as functions of z. The assumptions made when calculating
these functions were that the cell gap is 10 um, the liquid-crystal is MBBA-doped to give positive dielectric
anisotropy, the ordinary and extraordinary refractive indices for helium-neon laser light (633 nm) are

(5.55)
Figure 5.14 Tilt (0) and turn ( ) angles of the TN cell as functions of z 12 . Numbers in the figure correspond to
voltages normalized to Vth : V/Vth =(1) 1, (2) 1.083, (3) 1.295., (4) 1.69, (5) 2.56, (6) 3.42, (7) 4.12.

(5.56)

and the glass and the ordinary axis of the liquid-crystal have matching refractive indices. It was also assumed that k 11
/k 33=0.79 and k 22/k 33 =0.48.

Figure 5.15 shows the computed transmittance as a function of angle of incidence for various field strengths. In this
example, the polarizer transmits optical electric vectors along the y-axis (the director at the bottom is along the x-axis)
and the analyzer is parallel to the polarizer. This configuration differs from the commonly used one in which the
polarizer transmits optical electric vectors along the x-axis. But because curves for the case when the polarizer is
oriented parallel rather than perpendicular to the x-axis differ only about one percent from the curves shown in Figure
5.15, these results can be extended to the commonly used TN cell configuration.
Figure 5.15 Transmittance as a function of the angle of incidence in the x-z plane of Figure 5.6 12 . Curves are for the
seven values of V/Vth given in Figure 5.14.

Figure 5.16 shows the transmittance curves corresponding to the case where the direction of the incident beam around
the z-axis is turned 45° around the z-axis. The cell and the polarizers are unaltered. The difference between Figures
5.15 and 5.16 is clearly seen and the viewing angles are broader for the case of Figure 5.16. The cell configuration for
which the curves shown in Figure 5.16 were calculated is therefore usually used in a practical TFT/LCD (see Figure
2.2). Although there is a difference between the model described here and the practical display, the data in Figure 5.16
are good measures for describing the behavior of TFT/LCD viewing-angle characteristics in the vertical direction.

To make a twisted-nematic cell, the alignment or the director orientation at the glass substrate has to be controlled
precisely. The polyimide thin-film coated on the glass is rubbed in a particular direction. The rubbing produces grooves
in the polyimide film and these grooves make the liquid-crystal molecules align in the rubbing direction. Since the
liquid-crystal is rod-like, the energy difference between the alignments
Figure 5.16 Transmittance as in Figure 5.15 except that the turn angle 0 at the bottom is 45°, or the incident light is
turned -45° about the z-axis with respect to the configuration for Figure 5.15 12 . Curve labels are again as in Figure
5.14.

parallel to and perpendicular to the groove becomes large 2 . Strong anchoring or planar alignment thus results, and the
pretilt angle (the angle between the glass substrate and the director) becomes very small.

From a practical viewpoint, however, strong anchoring is not a good choice. The small pretilt angle (0(0)=0(d)=0)
tends to induce a reverse tilt, and when the reverse tilt occurs the cell is divided into domains with different twist
directions. Between the domains with right-hand twist and the left-hand twist there appears a line corresponding to the
domain boundary. This boundary is called a disclination line and gives the liquid - crystal a characteristic texture often
called "Schlieren textur" or "structure a noyaux." This disclination line is analogous to the dislocation in the crystalline
solid and has to be eliminated to prevent degradation of the picture quality. The domain structure in the liquid-crystal
is similar to the polycrystalline structure in the solid. The polycrystalline structure, however, can be eliminated in the
nematic liquid-crystal cell if the sample is prepared properly. The liquid-crystal in the TFT/LCDs has a single-
crystalline nature over the whole display. This is a very important characteristic of the nematic phase.

To eliminate the formation of domains or the occurrence of the reverse tilt, several methods have been proposed and
devised. The most popular method is to use a finite pretilt angle rather than a planar strong anchoring. A few to several
degrees of pretilt angle are generally used in the practical display. The surface treatment of the rubbing scheme and the
material of the alignment layer are chosen to set the angle of the pretilt.

As described above, liquid-crystal displays have an inherent drawback in that their viewing-angle characteristics are
not as good as those of CRTs, plasma-displays, and electroluminescent displays. This is because the birefringent effect
depends on the path-length of the cell and the path-length depends on the viewing angle of the display. To relax the
strong dependence of the transmission on the viewing angles, the alignment scheme shown in Figure 5.17 is usually
used (from discussion of Figure 5.16). The optical properties of TN cells correspond to the case of the light incidence
perpendicular to the panel. If the panel is rotated vertically or horizontally, the effective path-length or u increases and
this gives rise to the change of transmission. Figure 5.18 shows an example of the transmittance variation, for 16 gray
scales, as a function of the viewing angle when the panel is rotated vertically 13 . The curves show an asymmetrical
behavior with respect to the normal
Figure 5.17 Alignment scheme usually used in TFT/LCDs. The alignment direction is turned 45° at the bottom glass
interface and is turned -45° at the top glass interface.

view. The range of the viewing angle where the 16 gray scales can be recognized is not so wide, ranging from +5° to -
20°.

To improve the viewing-angle characteristics shown in Figure 5.18, a two-domain TN cell scheme has been proposed
14 . This scheme is

Figure 5.18 Angular dependence of the transmission when a TFT/LCD is rotated vertically 13 . Transmittance of 16
gray scales are shown as functions of the viewing angle.
Figure 5.19 Rubbing directions and pretilt angles of two-domain twisted-nematic liquid-crystal displays 14 . This
scheme improves the viewing-angle characteristics.

shown in Figure 5.19. The cell has two regions with different alignments, and the rubbing direction of region 1 in
Figure 5.19 differs from that of region 2 so that the pretilt angles are opposite. In this configuration the characteristics
of the two domains are summed and averaged; and as a result, both symmetrical transmittance with respect to the
normal incidence and wide viewing-angle characteristics are obtained. Figure 5.20 shows another scheme that makes
use of the difference in the pretilt angles in two regions 15 . These regions use different materials for the alignment
layer: organic and inorganic materials. The pretilt
Figure 5.20 Domain-divided twisted-nematic liquid-crystal cell 15 . Low- and high-pretilt angles are combined to
produce wide viewing-angle characteristics. The alignment layers are organic (shaded area) and inorganic.
Figure 5.21 Angular dependence of the transmission along the vertical direction of the TN cell. Symmetrical behavior
with respect to the normal incidence was obtained by a combination of different rubbing directions and low and high
pretilt angles. At the bottom, rubbing directions are different in two regions and pretilt angles are high. At the top,
rubbing directions are uniform and pretilt angles are low 13 .

angles are therefore different even though the rubbing direction is the same in both regions. Figure 5.21 shows
transmittance vs. viewing angle 13 for a cell combining the schemes shown in Figure 5.19 and 5.20. The symmetrical
behavior with respect to the normal incidence and the wide viewing-angle properties are obvious.

5.4. Super-Twisted Nematic (STN) Cell


The transmission characteristics of the twisted-nematic cell shows a moderate transition from on- (off-) to off- (on-)
state above the threshold voltage. This transition has a linear feature and the gray-scale representation is produced
relatively easily, especially in TFT/LCDs. In the passive-matrix liquid-crystal display, however, this slow transition
becomes a liability rather than an asset. The multiplexing in the passive-matrix LCDs with large information capacities
requires a fast transition from the off-state to the on-state above the threshold. Otherwise, the contrast ratio of these
displays becomes low and the viewing angle becomes narrow when the rows and columns of passive-matrix liquid-
crystal displays are highly multiplexed. According to Alt and Pleshko 16 , a matrix with m rows is driven optimally
when the applied rms voltages are chosen as follows:

(5.57)

and

(5.58)
where V s and V ns are respectively the rms voltages to the select pixel and to the non-select pixel 17 . As m, the
number of scan lines, is increased, the select voltage approaches the non-select voltage of Equation (5.58). If m=100,
the select voltage becomes Vs =1.11. If m=400, V s =1.05. Therefore, the select voltage is higher than the non-select
voltage by only a few percent.

Super-twisted nematic (STN) liquid-crystal displays were developed to satisfy the requirements described by
Equations (5.57) and (5.58). The STN is a cell with a twist angle of about 270° and with a relatively high pretilt angle.
The phenomenon produced with this kind of cell was originally referred to as the supertwisted birefringent effect
(SBE) 18 . The original SBE technology has been modified (lower pretilt angles, etc.) and the term STN (super-twisted
nematic) 19 , 20 is now used most commonly. The basic principle of an STN cell, however, is the same as that of an
SBE cell. Figure 5.22 shows the voltage dependence of the director in the midplane of a chiral nematic layer with a
28° pretilt angle at both boundaries. The bistable range appears as the total twist angle is increased above 245°, and
the technologically convenient twist angle of 270° is generally used.

A high pretilt angle, on the order of 5° and 30°, is required at both interfaces to ensure that only deformations of a
twist angle of about

Figure 5.22 Theoretical curves of the tilt angle of local directors in the midplane of an STN cell as a function of the
reduced voltage V/Vth where Vth is the Freedericksz threshold voltage of a non-twisted layer with zero pretilt 18 .

270° occur in the display. With low pretilt angles, a distortion with 180° or less twist becomes more stable. The
polarizer setting of STN cells also differs from that of TN cells. For a nematic layer with a 270° left-handed twist, the
optimum state is obtained when the front polarizer is oriented so that (1) the plane of vibration of the E vector makes a
30° angle with the projection of the layer and (2) the rear polarizer is at an angle of about 60° with respect to the
projection of the director at the rear boundary. This orientation is required because of the residual twist and retardation
of the select state. As a result of the interference of the optical normal modes propagating in the layer, the display has
a yellow birefringence color in the non-select state. Rotation of one of the polarizers by an angle of 90° results in an
image, complementary to the previous "yellow mode," in which the select state is colorless and bright and the non-
select state is dark blue ("blue mode"). Recent development of the retardation film has made the "white mode" STN
display possible.

The threshold voltage of the STN display is given by 21


(5.59)

where is the total twist angle, 0 0 is the pretilt angle, p is the chiral pitch, and d is the cell spacing. The threshold
voltage is 2–3 V, and the response time of the STN cell is on the order of a few hundred ms at 20°C. A contrast ratio
of 10:1 or more can be obtained and the viewing cone makes an angle of 30° with the vertical.

References
1. Schadt, M., and Helfrich, W. (1971). Voltage-dependent optical activity of a twisted nematic liquid-crystal. Applied
Physics Letters , 18, 127–128.

2. Nehring, J., Kmetz, A.R., and Scheffer, T.J. (1976). Analysis of weak-boundary-coupling effects in liquid-crystal
displays. Journal of Applied Physics , 47, 850–857.

3. Freedericksz, V., and Zolina, V. (1933). Trans. Faraday Society , 29, 919.

4. Gruler, H., Scheffer, T.J., and Meier, G. (1972). Elastic constants of nematic liquid-crystals. Z. Naturforsch ., A27,
966–976.

5. Ohtsuka, T. (1991). Basic theory and physical characteristics of liquid-crystal. In Liquid Crystalline Materials ,
edited by S. Kusabayashi, p. 33. Tokyo: Kohdansha (In Japanese).

6. Jakeman, E. and Raynes, E.P. (1972). Electro-optic response times in liquid-crystals, Physics Letters , 39A, 69–70.

7. Katoh, K., Imagi, S., and Kobayashi, N. (1988). Active-matrix-addressed color LCDs for avionic application. In
Digest of Technical Papers of the Society for Information Display International Symposium (Anaheim, 1988), pp. 238–
241. California: SID.

8. Gooch, C.H. and Tarry, H.A. (1975). The optical properties of twisted nematic liquid-crystal structures with twist
angles 90°. Journal of Physics D: Applied Physics , 8, 1575–1584.

9. Gooch, C.H., and Tarry, H.A. (1974). Optical characteristics of twisted nematic liquid-crystal films. Electronics
Letters , 10, 2–4.

10. Azzam, R.M.A. and Bashara, N.M. (1972). Simplified approach to the propagation of polarized light in anisotropic
media-application to liquidcrystals. Journal of the Optical Society of America , 62, 1252–1257.

11. Funada, F., Okada, M., Kimura, N., and Awane, K. (1988). Selection and optimizing of liquid-crystal display
modes for the full color active-matrix LCDs, Journal of the Institute of Television Engineers , 42, 1029–1034 (In
Japanese).

12. Berreman, D.W. (1973). Optics in smoothly varying anisotropic planar structures: application to liquid-crystal twist
cells. Journal of the Optical Society of America , 63, 1374–1380.

13. Takatori, K., Sumiyoshi, K., Hirai, Y., and Kaneko, S. (1992). A complementary TN LCD with wide-viewing-
angle grayscale. In Proc. of the 12th International Display Research Conference (Hiroshima, 1992), pp. 591–594.
California: SID, Tokyo: ITE.

14. Yang, K.H. (1991). Two-domain twisted nematic and tilted homeotropic liquid-crystal displays for active matrix
applications. In Proc. International Display Research Conference (San Diego, 1991), pp. 68–72, California: SID.

15. Koike, Y., Kamada, T., Okamoto, K., Ohashi, M., Tomita, I., and Okabe, M. (1992). A full-color TFT-LCD with a
domain-divided twisted-nematic structure. In Digest of Technical Papers of the Society for Information Display
International Symposium (Boston, 1992), pp. 798–801. California: SID.

16. Alt, P.M., and Pleshko, P. (1979). Scanning limitations of liquid-crystal displays. IEEE Transactions on Electron
Devices , ED-21, 146–155.

17. Nehring, J., and Kmetz, A.R. (1979). Ultimate limits for matrix addressing of rms-responding liquid-crystal
displays. IEEE Transactions on Electron Devices , ED-26, 795–802.

18. Scheffer, T.J., and Nehring, J. (1984). A new, highly multiplexable liquid-crystal display. Applied Physics Letters ,
45, 1021–1023.

19. Leenhouts, F., and Schadt, M. (1986). Electro-optics of supertwist displays: dependence on liquid-crystal material
parameters. In Proc. 6th International Display Research Conference (Tokyo, 1986), pp. 388–391. California: SID,
Tokyo: ITE.

20. Kinugawa, K., Kondo, Y., Kanasaki, M., Kawakami, H., and Kaneko, E. (1986). 640×480 pixel LCD using highly
twisted birefringence effect with low pretilt angle. In Digest of Technical Papers of the Society for Information
Display International Symposium (San Diego, 1986), pp. 122–125. California: SID.

21. Breddels, P.A., and van Spraing, H.A. (1985). An analytical expression for the optical threshold in highly twisted
nematic systems with nonzero tilt angels at the boundaries. Journal of Applied Physics , 58, 2162–2166.

CHAPTER 6
Liquid-Crystal
In a crystal state of matter, the solid forms a three-dimensional lattice having long-range order. When it is heated
above the melting point, it turns to an isotropic liquid having neither long-range nor short-range order. The liquid-
crystal is an intermediate state of matter between a solid crystal and an isotropic liquid (Figure 6.1).

Material showing a liquid-crystalline phase is composed of many rod-like molecules. Due to this molecular feature or
anisotropy, the solid does not immediately turn into an isotropic liquid at the melting point, T m . When this material
reaches T m , the solid changes into a transitional liquid-crystalline phase. In this phase, the gravitational or positional
order of constituent molecules is lost as in a normal liquid. However, there still remains some degree of orientational
order of molecules. If this material is further heated above the clearing point, T c, this fluid turns into an isotropic
liquid. The intermediate state between these two phase transitions is a liquid-crystalline phase. In this phase, the fluid
appears turbid and is found to be strongly birefringent when observed between crossed polarizers. This phase is
sometimes called the mesophase, since there is a contradictory tone in the name "liquid-crystal".

Figure 6.1 The liquid crystalline phase is defined as an intermediate state between crystal and liquid. The thermotropic
liquid crystalline phase appears in the temperature range between the melting point, Tm , and the clearing point, Tc.

This type of liquid-crystal, in which the mesophase is defined by the temperature range between T m and T c, is called
thermotropic. In another type of liquid-crystal, called lyotropic, the amount of solvent defines the mesophase. Most
liquid-crystals used for display applications are thermotropic. Apart from the thermotropic and lyotropic definitions,
liquid-crystal can be classified into three types, i.e., nematic, cholesteric, and smectic liquid-crystals (G. Friedel, 1922
1 ).

Among these three types, the nematic phase is most widely used. The two-dimensional molecular alignment of nematic
liquid-crystal is shown schematically in Figure 6.2. The molecules in the nematic liquid-crystal are aligned with their
long axes nearly parallel to each other. Though they have no long-range correlation between their centers of mass, they
have an orientational order or a local preferred direction. The orientational axis is generally described by a unit vector
n , the director.

Although the rod-like molecules of liquid-crystal change their directions quite rapidly, fluctuations in the director n are
relatively long range. The period of spatial fluctuations in n is nearly equal to the wavelength of visible light. This
results in strong light scattering, and the liquid-crystal appears as a turbid fluid. Each molecule deviates from the
average direction n by an angle of 0. The director has no polarity, in other words, n and - n are equivalent. This
equivalence is, of course, a macroscopic effect. The molecules in the liquid-crystal are not symmetrical in their
configuration and have distinct heads and tails. However, these heads and tails are equal in number. This is why n and
- n are equivalent.

Microscopic observation of nematic liquid-crystal shows the texture patterns, or threads, which correspond to
discontinuities in the local preferred direction of molecules (the word "nematic" comes from a Greek word meaning
thread). One classical example of nematic liquid-crystal is p-azoxyanisole (PAA):

(6.a)

The two benzene rings are nearly coplanar, and roughly speaking, this is a rigid rod about 20 Å long and 5 Å wide.
Another classical example is N-(p-methoxybenzylidene)-p-butylaniline (MBBA):
Figure 6.2 Two-dimensional schematic representation of the orientation of nematic liquid-crystal molecules. Each rod-
like molecule fluctuates quite rapidly, but has a definite orientational order expressed by a unit vector n, the director.
The angle between the orientation of the director and each molecule is 0.

(6.b)

A broad class of organic molecules with the same general pattern

(6.c)

also gives mesophases. In the center of the molecule, there is a core with two terminals, one at each end, i.e., R and R'.
The core is composed of two benzene rings with a connector in between. The combination of a variety of terminals
and connectors, X, is known to yield a nematic-phase liquid-crystal. In some cases, the benzene rings are replaced by
other ring structures like pyridine, pyrimidine, or dioxane.

These still exhibit the nematic phase, however, it is not possible at present to specify which structures will become a
liquid-crystal. When the connector between two benzenes is azoxy, and the terminals are two methoxy bases, this
organic material (PAA) exhibits the nematic phase in a rather high temperature range between 117°C and 136°C. This
liquid-crystal material had been studied quite extensively until H. Kelker 2 introduced the first room-temperature
nematic liquid-crystal in 1969 (MBBA). This organic compound's connector is a Schiff base, and the two terminals are
a methoxy base and a butyl base. It exhibits the nematic phase in the temperature range between 22°C and 47°C.
When two benzene rings are connected directly (biphenyl core) and the two terminals are a cyano base and a pentyl
base, it is cyano-pentyl-biphenyl (6.d), which also exhibits the nematic phase between 24°C and 35°C, and is widely
used in display applications due to its chemical stability and large anisotropic dielectric constant. When the connector
is also benzene, it has a core of terphenyl. Some other examples of popular biphenyl- and terphenyl-core liquid-
crystals are shown below.

(6.d)

(6.e)

(6.f)

(6.g)

The cholesteric phase of liquid-crystals is similar to the nematic phase. However, in this phase there is spatial variation
in the director along the helical axis, as shown in Figure 6.3. In each plane of this structure, the orientation is the same
as in the nematic liquid-crystal. If the helical axis is taken as the z-axis, the director of the cholesteric phase can be
expressed as follows:

(6.1)

(6.2)

(6.3)

Here, p is the full rotation period of the director, and the phase constant is assumed to be zero. Since n and -n are
equivalent, the repetition period in the cholesteric phase is p/2. In other words, the cholesteric phase has a layered
structure with a period of p/2. The extraordinary axis of the cholesteric liquid-crystal is parallel to the helical axis, and
this means that the cholesteric liquid-crystal is optically negative (n e <n 0 ). The name "cholesteric" comes from the
fact that many derivatives of cholesterol exist in this phase. An alternative name is the chiral nematic phase since many
liquid-crystals with a chiral carbon have the same characteristics.

The last of the three phases of liquid-crystals is the smectic phase. This exhibits a much more viscous nature than the
nematic phase. This is because this mesophase has a layered structure, as shown in Figure 6.4. The distance between
the neighboring layers is about the length of the long molecular axis, which is of the order of a few angstroms. The
existence of this layered structure is directly evidenced by x-ray diffraction patterns, which correspond to layer
thickness. In each layer, the liquid-crystal molecules are aligned perpendicular to the layer (smectic A or SA ). The
orientational order is similar to that in the nematic phase and the molecular centers of gravity have no long-range
order. This smectic-A phase is the simplest case of the smectic phase. In the smectic-C phase, the molecules are not
normal to the layer, but are somewhat tilted. There are other types of smectic phases in addition to SA and SC, that
exhibit some crystalline nature such as pseudo-hexagonal packing or herring-bone-type packing in the layers.

Figure 6.3 Molecular alignment of a cholesteric liquid crystal. In this phase, the director changes its orientation
gradually along the helical axis. This helical structure, with its full period of p, is characteristic of the cholesteric
phase, and the helical axis coincides with the optical axis of this material.
Figure 6.4 Schematic representation of the smectic phase of liquid-crystal. This phase has a layered structure in which
the molecular orientation is perpendicular or nearly perpendicular to the layer. This figure shows an ordered layer
structure. However, the molecular fluctuations are very similar to those in other liquid-crystalline phases.

As mentioned above, the director n is defined as the average alignment direction of the liquid-crystal molecules. In the
Cartesian coordinate system shown in Figure 6.5, this is taken as the z-axis. Qualitatively, the nematic liquid-crystal is
said to be "more ordered" than the isotropic liquid. In many instances, however, it is more convenient to view this on a
quantitative basis. We need to define an order parameter which expresses the state of alignment of rod-like molecules.
Figure 6.5 Cartesian coordinates used to derive the order parameter of the liquid-crystal. The director orientation is
parallel to the z-axis. The polar angle 0 is different from the one defined in Figure 5.6.

If each rod is expressed by a unit vector a , and assumed to have complete cylindrical symmetry, the Cartesian
component will be given by its polar angles 0 and , as shown in Figure 6.5:

(6.4)

(6.5)

(6.6)

Note that the definition of 0 here is different from that given in Figure 5.6.

In developing some measure of the alignment in the nematic phase, the state of rod alignment may be described by a
distribution function f(0, )dn which gives the probability of finding molecules in a solid angle dn=sin0 d0 d .
Although the distribution function f(0, ) cannot be determined completely, some general features of f(0, ) can be
discussed. First, since the nematic phase has complete symmetry about n , f(0, ) is independent of . Secondly since
there is no directional preference ( n and -n are equivalent), f(0)=f( -0). Therefore, f(0) will have a strong peak
around 0=0 (or ) and will have a minimum at 0= /2. Since our target is not to describe the distribution function
in full detail, but to have some numerical parameter, the most logical first step in
Figure 6.6 Temperature dependence of the order parameter S of the nematic liquid-crystal (PAA) 4 . The curve
represents the average of data obtained from optical, x-ray, and NMR experiments. The experimental data are well
described by theoretical calculation.

characterizing the alignment would be to use a statistical average denoted by angular brackets:

(6.7)

However, this average would be zero due to the property f(0)=f(0- ) and is therefore not adequate for our purpose.
Thus, we proceed to the next step.

The next option is cos 2 0 and the order parameter S, first introduced by Tsvetkov in 1942 3 , is given by

(6.8)

If all the molecules are aligned parallel to the director, the order parameter S is equal to 1. If, on the other hand, all the
rods are perpendicular to n , S is equal to -1/2, although it is unlikely that the rod-like molecules are aligned
perpendicular to one another. When the molecular orientation is entirely random, as in an isotropic liquid, we would
have <cos 2 0>=1/3 and the order parameter S would be zero. For practical nematic liquid-crystals, S takes values from
0.3–0.4 at T c (T NI) to 0.6–0.8 at lower temperatures. An experimental example of the temperature dependence of the
order parameter is shown in Figure 6.6 4 . The discussion of order parameter can be extended to that of liquid-
crystalline constants in terms of S. For example, the linear relation between the dielectric anisotropy and the order
parameter can be derived. However, we will not go further in discussing these relations, since its extendability is
somewhat limited.

In Chapter 5, we discussed the optical and electrical properties of TN cells which were found to depend on such
liquid-crystalline constants as the dielectric anisotropy, birefringence, elastic constants, and coefficients of viscosity. In
the following sections we will discuss the basic aspects of these constants.
6.1. Dielectric Permittivity and Refractive Index
The dielectric permittivity in a static electric field is easily obtained by measuring the capacitance when the plane
capacitor is filled with a specific material. The capacitance increases from vacuum dielectric permittivity o to r o
where r is the relative permittivity of the material. This increase in capacitance is a result of induced polarization P .
The relation between the electric field E and the electric displacement D is described as follows:

(6.9)

From Equation (6.9), the electric polarization is expressed as

(6.10)

This relation holds for any isotropic medium. However, in an anisotropic medium like liquid-crystal, the polarization is
modified from that in Equation (6.10) to a tensor formulation:

(6.11)

where X ij is an element of the susceptibility tensor X, and the subscript which appears twice means a summation on
that subscript.

Generally speaking, X is a tensor of rank two and this polarization tensor X ij is symmetric with respect to the two
subscripts, that is

(6.12)

Any tensor of rank two can be diagonalized by choosing an appropriate set of coordinates. Thus, the tensor can be
expressed as follows:

(6.13)

The relation X xx=X yy =X zz holds in isotropic media, such that

(6.14)

where d ij is a Kronecker delta. However, in an anisotropic medium like liquid-crystal, the susceptibility tensor is
expressed as

(6.15)

with the z-axis parallel to the director, n . Thus, the displacement vector D is given by

(6.16)

where - =(1+X xx) 0 and // = (1+Xzz ) 0. With the z-axis parallel to the director, the permittivity tensor is
diagonalized and is given by - and // , where the subscripts // and - are used for the directions parallel and
perpendicular to the director. In a nematic liquid-crystal, // is different from - . Therefore, generally speaking, the
displacement vector is not parallel to the electric field. The displacement vector becomes parallel to E only when the
electric field lies on the x-y plane or is parallel to E z .

Figure 6.7 Temperature dependence of the dielectric constant of cyano-biphenyl: C7 H15 CN at a frequency
of 100 kHz5 . This material exhibits a large positive anisotropy ( // > - ). The average of the dielectric constant av
= ( // +2 - )/3, denoted by a dotted line, is nearly equal to the isotropic dielectric constant above the clearing
temperature.

For display applications, a large positive value of /_\ (= // - - ) is preferable since it is directly related to the
threshold voltage of the liquid-crystal (Equations (5.12) and (5.25)). Many nematic crystals exhibit a positive /_\ ,
although some show negative values (MBBA for example). The dielectric permittivity ( // , - ) is dependent on
temperature, as shown in Figure 6.7 5 . It also depends on the frequency of electric excitation. In measuring the plane
capacitor filled with nematic liquid-crystal, the ac frequency is set low enough (1 kHz to 10 kHz) to yield static
permittivities. In the optical frequency range, the dielectric permittivity is obtained from the refractive index n as
shown below:

(6.17)

(6.18)

The liquid-crystal exhibits anisotropy of refractive indices or bire-fringence, as do optically uniaxial crystals. A
uniaxial crystal has two principal refractive indices n o and n e : n o is the refractive index for an "ordinary" ray, in
which the electric vector of a light wave vibrates perpendicular to the optical axis and n e is the index for an
"extraordinary" ray, in which the vibration of the electric vector is parallel to the optical axis. The birefringence /_\n
corresponds to the difference between the principal refractive indices, n o and n e :

(6.19)

A liquid-crystal behaves like a uniaxial crystal. In the nematic phase, the optical axis coincides with the director of the
liquid-crystal. This leads to:

(6.20)

(6.21)

where the subscripts // and - correspond to directions parallel and perpendicular to the director. Thus, we have for the
birefringence

(6.22)

Generally speaking, the /_\n of the nematic liquid-crystal is positive. The temperature dependence of the refractive
indices, shown in Figure 6.8 6 , shows almost the same behavior as the dielectric permittivity. The dependence of the
refractive indices on wavelength is shown in Figure 6.9 7 . The average value of the refractive indices in the nematic
phase is given by the relationship

(6.23)

where the value is different from the refractive index n is in the isotropic phase, although this difference
is small. This is also true for the difference in the dielectric constants between the isotropic phase and nematic phase.
The average value of the dielectric constant av is calculated from the experimental values of // and - as
(6.24)

In a number of strongly positive materials (/_\ >0), the calculated mean dielectric constant av is a few percentage
points smaller than the isotropic value is. This has been confirmed in theoretical studies.

Figure 6.8 Temperature dependence of refractive indices n// and n - of MBBA at a wavelength of 546 nm 6 . Since the
dielectric permittivities in the optical frequency range are given by the square of the refractive indices, the dielectric
anisotropy is positive for the light wave. It is worth noting that the dielectric anisotropy of MBBA is negative in a
static field.

Figure 6.9 Refractive indices n// and n - as a function of wavelength. The liquid crystal is 4-butoxyphenylester of 4'-
hexyloxybenzoic acid 7 . This material shows a nematic phase in the temperature range between 50°C and 102°C. The
measurement temperature was 80°C.

In the cholesteric phase of liquid-crystals, shown in Figure 6.3, the optical axis becomes the helical axis, which is
perpendicular to the director. Thus, for cholesteric material, the following relations hold:
(6.25)

(6.26)

Here, the relation n // >n - still holds, meaning that n e is smaller than n o . Consequently, in the cholesteric liquid-
crystal the birefringence becomes negative.

The dielectric permittivities and refractive indices of PAA and MBBA are listed in Table 6.1 8 along with other
physical constants.

6.2. Elastic Constants


In uniaxial liquid-crystals, the preferred orientation of liquid-crystal molecules is given by the director n . This
orientation is imposed by a surface treatment at the boundaries or by an external field. The nematic liquid-crystal is
composed of rod-like molecules of which the long axes are aligned nearly parallel to those of neighboring molecules.
This preferred orientation of neighboring molecules at each point r is given by the vector n ( r ). The orientation varies
gradually and continuously from point to point within the medium. These changes take place over a macroscopic
distance (a few microns). The orientation is usually imposed at the boundaries between the liquid-crystal and the
alignment layers deposited on glass substrates.

The anchoring direction of molecules perpendicular to the planar boundary is called homeotropic while the direction of
molecules parallel to the surface is called homogeneous or planar. This anchoring has to be strong enough not to be
affected by external forces or fields. However, even with strong anchoring, the liquid-crystal molecules separated from
the boundaries by a distance larger than the coherence length change their directions quite easily with the application
of an external field.

This transition of the director from one direction to the other induces a curvature strain in the medium. By introducing
a local right-handed system of Cartesian coordinates x, y, z with n parallel to the z-axis at the origin, we can define the
curvature strain tensors n ij (n ij =dn i/d xj ), which correspond to the changes in the orientation of the director.
Ignoring dnz /dx, dnz /dy, and dnZ /dz ( n 2 =1), we obtain
(6.27)

The six components of curvature strain are


(6.28)

These three types of deformation, splay (s2 ), twist (t1 ), and bend (b 2 ), are shown in Figure 6.10. For these
deformations in an incompressible fluid and for isothermal deformation, the free energy per unit volume F is given as

(6.29)

For infinitesimal deformation, this F value of a deformed liquid-crystal relative to that in the state of uniform
orientation, /_\F, can be expanded as

(6.30)

where k ij is an elastic constant and a tensor of rank two, generally expressed as follows:

(6.31)

The number of components can be reduced due to the presence of cylindrical symmetry:

(6.32)

where k 1 and k 2 are abbreviations for k 11 and k 12 . In the nematic liquid-crystal, this is further simplified since the
nematic phase is non-polar and non-enantiomorphic. Each molecule of the nematic liquid-crystal
Figure 6.10 Deformation of nematic liquid-crystal molecules. Three types of deformation, splay, twist, and bend, are
shown for variations in the director orientation along the y-axis.

may be polar, but there is equal probability that they may point in either direction. Consequently the nematic liquid-
crystal is non-polar (k 1 =0). Moreover, in non-enantiomorphic material, the tensor component k 2 is equal to zero.
Thus, both k 1 and k 2 in Equation (6.32) are equal to zero.

Therefore, we must go further in discussing the curvature strain. The free-energy density is expanded as follows:

(6.33)

The tensor k ijlm generally has 81 components, where 1 i, j, l, m 3. This notation will be abbreviated here as kij,
where 1 i, j 9. Two examples are k 1111 k 11, and k 3333 k 99. Since dnz /d x, dnz /d y, and dn z /d z are
neglected, the number of components is reduced from 81 to 36. Cylindrical symmetry reduces this number to 18 with
only five independent constants. In the nematic phase, this is further reduced to four constants.
(6.34)

where k 15=k 11-k 22-k 24. The free-energy density of deformation is now written as

(6.35)

In equilibrium situations, we can omit k 24 and in a planar configuration the last term can also be omitted, since we
can write

(6.36)

Equation (6.35) can be given in vector notation, such that

(6.37)

(6.38)

(6.39)

From Equations (6.35), (6.37), (6.38) and (6.39), we obtain the fundamental expression for the free-energy density for
nematics:

(6.40)

The three elastic constants k 11 , k 22, and k 33 are called the splay, twist, and bend constants, respectively. These
constants correspond to the three types of deformation shown in Figure 6.10 and are important in determining the
optical and electrical properties of a twisted nematic cell.

There is no good way to directly measure the elastic constants. However, they can be estimated if other constants like
the magnetic susceptibility are known. The temperature dependence of the elastic constants of nematic materials is
shown in Figure 6.11 9 . The data in this figure show the average of somewhat scattered data. This is unavoidable due
to the difficulties involved in achieving reproducible surface treatment during sample preparation. The set of physical
constants for the nematic liquid-crystals in Table 6.2 should also be read with these limitations in mind.

Light scattering by the liquid-crystal can be explained in terms of these elastic constants. The experimental results
regarding light

Figure 6.11 Temperature dependence of the elastic constants of nematic liquid crystals: (a) PAA, (b) MBBA 9 . /_\X is
the anisotropy of the magnetic susceptibility (/_\X=X // -X - : dimensionless quantity). Absolute values are listed in
Table 6.2.

scattering indicate that scattering is quite strong for small scattering vector q . Moreover, scattering is intense for
crossed polarization, i.e., if the polarizations of the incoming and outgoing lights are not parallel. The molecules of the
liquid-crystal fluctuate from their equilibrium orientation, of which the director is given as n o . Thus, the director for
the liquid-crystal at r is written as

(6.41)

where dn(r) is a small fluctuation in n . Since dn(r) is perpendicular to n o , the x and y components of dn(r) , n x and
n y , are measures of molecular fluctuation and light scattering. To discuss the scattering due to the director fluctuation,
we start from Equation (6.40) giving the total curvature free energy for the nematic phase:

(6.42)

The scattering of light is characterized by the wave vectors k i and k o of the incoming and outgoing beam. The
difference,

(6.43)

is the scattering vector, and the fluctuations n x ( r ) and n y ( r ) are expanded in Fourier series as

(6.44)

and

(6.45)

The amplitude of the wave vector is given by

(6.46)

where n is the refractive index of the material and is the wavelength of light. Since the liquid-crystal is anisotropic,
the amplitude of the wave vector depends on the angle of incidence and the director orientation.

From Equations (6.44) and (6.45), we obtain relations like

(6.47)

where

(6.48)

To calculate the free energy of Equation (6.42) using Equation (6.47), it is helpful to rotate the coordinate system x, y,
z for each q around the z-axis, so that the vector q lies on the new plane (x', z). In this new coordinate system (x', y', z),
q has no y' component (q y =0). Thus, the free energy takes a simpler form:
(6.49)

where n 1 (q) and n 2 (q) are linear transformations from n x (q) and n y (q) . n 1 (q) and n 2 (q) lie on the (x', z) plane
and (y', z) plane, respectively, q// =q cos0 and q - =q sin0 where 0 is the angle between the vector q and the z-axis.

We can now estimate the thermal average of |n 1 (q) | 2 or <|n 1 (q) | 2 >. For this, we use the equipartition theorem. For
a system in thermal equilibrium, the average energy per degree of freedom is equal to (1/2) kT. Therefore, we obtain

(6.50)

(6.51)

This is the main equation of fluctuation theory first described by de Gennes. These fluctuations in n give rise to
fluctuations in optical dielectric permittivity, and thus to light scattering. The scattering intensity is represented by the
differential cross-section (per unit solid angle of the outgoing beam around the direction k 0 ) and is shown to be
proportional to the sum of <|n a (q) | 2 > (a=1, 2). This sum is large for small values of q. The wavelength of visible
light is far higher than the molecular dimensions, so q becomes small, resulting in a large disturbance and strong
scattering. A rough estimate from these discussions leads to the conclusion that scattering from the liquid-crystal is
stronger than that from a usual liquid by a factor of 106 . This is the reason the liquid-crystal appears turbid in a glass
tube.

6.3. Viscosity
Nematic liquid-crystal flows like a conventional organic liquid having similar molecules. However, its behavior is
rather complex and therefore difficult to study due to the anisotropy of nematics. The translational motions of
molecules are coupled to inner, orientational motions and the flow is dependent on the angles the director forms with
the flow direction and with the velocity gradient. Thus, in most cases, the flow disturbs the alignment and causes the
director to rotate. Conversely, rotation of the director induces back flow in the surrounding nematics. As a result, the
theoretical and experimental studies are more complex than those for isotropic liquids.

Before going on to a discussion of liquid-crystals, we should first consider the case of an ideal liquid with no viscosity.
The state of a fluid is described by its velocity v, pressure p, and density p. The equation of continuity is given by

(6.52)

If we assume that the fluid is incompressible (which holds for all practical purposes),

(6.53)

Equation (6.52) becomes


(6.54)

Next, we should consider the equation of motion:

(6.55)

where f is the force per unit volume. Since the contribution to acceleration dv/dt comes from the time derivative and
the velocity gradient, and the contribution to total force comes from the pressure gradient (other external forces are
neglected), Equation (6.55) is reduced to

(6.56)

where p is the pressure.

Now, we must consider a viscous fluid. The viscosity term f visc is added to Equation (6.56) and the equation of
motion is written as

(6.57)

We define the stress tensor 6 here as

(6.58)

where 6 ij , is an element of the stress tensor, d ij is a Kronecker delta, and is an element of the viscous-stress
tensor. Thus, Equation (6.57) is reduced to

(6.59)

where x j (j=1, 2, and 3) corresponds to x, y, and z, and v i (i=1, 2, and 3) corresponds to v x, v y, and vz. When the
velocity gradient is small, the viscous-stress tensor may be assumed to be linear in the spatial derivatives of velocity,
dv i /dx j :

(6.60)

where n is the coefficient of viscosity measured in poise (g s -1 cm- 1 ).

The stress tensor is proportional to the symmetric part of the velocity-gradient tensor: dv i /dx j . This symmetric part
is often referred to as tensor A, of which an element is given by

(6.61)
From Equation (6.54), the trace of tensor A is expected to vanish for incompressible fluids. The asymmetric part of the
velocity-gradient tensor is written as tensor W , of which an element is given by

(6.62)

Figure 6.12 A simple shear flow is decomposed into an irrotational flow and a rotational flow.

which is related to the local angular velocity of the fluid. The decomposition of the velocity-gradient tensor into a
symmetric and an asymmetric part means that the fluid flow can be described as a superposition of an irrotational flow
( W =0) and a rotational flow ( A =0). This decomposition and superposition are illustrated in Figure 6.12.

Up to now, we have treated the motion of fluid in its general form. In nematic liquid-crystals, however, the equation
for motion of the director has to be taken into account. According to an analysis of coupling between orientation and
the flow of nematic molecules 10 , 11 , the viscous (hydrodynamic) part of the stress tensor, which is the total stress
minus the static (or elastic) part, can be described as a function of the velocity-gradient tensor A and the angular
velocity N of the director n relative to its surroundings. The element N is written as

(6.63)

where W ij is given by Equation (6.62). Since the stress tensor in Equation (6.60) must be an even function of n , the
most general form for an incompressible nematic is given by

(6.64)

The six coefficients, from a 1 to a 6 , have the dimensions of viscosity and are called Leslie coefficients 10 . However,
only five coefficients are independent since there is one relation, derived by Parodi 11 .
Figure 6.13 The coordinate system used to analyze the simple shear flow case. The orientation of the director is fixed
by external forces like an electric field.

(6.65)

To consider the viscous-stress tensor more concretely, we refer to a simple shear flow in a nematic liquid-crystal.
Under the assumption that the flow runs along the z-axis between two parallel plates normal to the x-axis, which is the
direction of the velocity gradient (Figure 6.13), the viscosity stress given in Equation (6.60) is reduced to

(6.66)

where n is the viscosity coefficient, and u is the velocity written as v = [0, 0, u(x)].

With a simple shear flow, as shown in Figure 6.13, where the velocity gradient is constant, the flow is laminar. The
force required to keep this laminar flow is the stress given in Equation (6.66). This stress depends on the orientation of
the director n . Figure 6.14 shows three limiting cases:

(a) n 1 : the director is parallel to the x-axis (velocity gradient)

(b) n 2 : the director is parallel to the z-axis (flow direction)

(c) n 3 : the director is parallel to the y-axis (normal to the shear plane).

The coefficients n 1 , n 2 , and n 3 are called Miesowicz coefficients. The Miesowicz's viscosity coefficients for MBBA
are shown in Figure 6.15 as a function of temperature 12 . For a fixed director at arbitrary angles 0 and (Figure
6.13), the effective viscosity is given by

Figure 6.14 Miesowicz's viscosity coefficients are defined for the three limiting cases of director orientation with
respect to the shear plane.

Figure 6.15 Viscosity coefficients n 1 , n 2 , and n 3 of MBBA as a function of temperature 12 . The temperature scale is
linear in 1/T. The coefficient n 3 behaves very much like that in a continuation of isotropic viscosity.

(6.67)

where n 12 is shear stress corresponding to a stretch type of deformation which is depicted in Figure 6.14 (d).

The relation between n and a is as follows:

(6.68)

(6.69)

(6.70)

(6.71)

As for the rotation of the director, the viscous torque is again assumed to be a linear function of the velocity gradients,
and of the motion of the director relative to its surroundings. The viscous torque C is thus written as
(6.72)

where Y 1 and Y 2 are the shear torque coefficients which have the dimensions of viscosity. These rotational viscosity
coefficients are related to the Leslie coefficients as follows:

(6.73)

(6.74)

The coefficients of viscosity Y i , a i , and n i are listed in Table 6.1 for PAA and MBBA.

References
1. Friedel, M.G. (1922). Ann. Physique , 18, 273.

2. Kelker, H. and Scheurle, B. (1969). Angew. Chem. , 81, 903; J. Phys. (Paris), Colloq., C-4 , 104.

3. Tsvetkov, W. (1942). Acta Physicochim. URSS , 16, 132.

4. Rowell, J.C., Phillips, W.D., Melby, L.R., and Panar, M. (1965). NMR studies of some liquid-crystal systems.
Journal of Chemical Physics , 43, 3442–3454.

5. Lippens, D., Parneix, J.P., and Chapoton, A. (1977). Journal de Physique , 38, 1465.

6. Blinov, L.M. (1983). Electro-Optical and Magneto-Optical Properties of Liquid Crystals , p. 43. Chichester: John
Wiley & Sons.

7. Blinov. L.M. (1983). Electro-Optical and Magneto-Optical Properties of Liquid Crystals , p. 44, Chichester: John
Wiley & Sons.

8. Blinov, L.M. (1983). Electro-Optical and Magneto-Optical Properties of Liquid Crystals , p. xxi, Chichester: John
Wiley & Sons.

9. De Jeu, W.H., Claassen, W.A.P., and Spruijt, A.M.J. (1976). The determination of the elastic constants of nematic
liquid-crystals. Molecular Crystals and Liquid Crystals , 37, 269–280.

10. Leslie, F.M. (1970). Distortion of twisted orientation patterns in liquid-crystals by magnetic field. Molecular
Crystals and Liquid Crystals , 12, 57–72.

11. Parodi, O. (1970). Stress tensor for a nematic liquid-crystal. J. Physique , 31, 581–584.

12. Gähwiller, C.H. (1973). Direct determination of the five independent viscosity coefficients of nematic liquid-
crystals. Molecular Crystals and Liquid Crystals , 20, 301–308.

General References
1. Amorphous and Liquid Semiconductors edited by J. Tauc. London. New York: Plenum Press (1974).
2. The Physics of Liquid Crystals by P.G. de Gennes. Oxford: Clarendon Press (1974).

3. Liquid Crystals by S. Chandrasekhar. Cambridge: Cambridge University Press (1977).

4. Electronic Processes in Non-Crystalline Materials by N.F. Mott and E.A. Davis. Oxford: Clarendon Press (1979).

5. Amorphous Semiconductors edited by M.H. Brodsky (Topics in Applied Physics Vol. 36). Berlin, Heidelberg, New
York: Springer-Verlag (1979).

6. Physical Properties of Liquid Crystalline Materials by W.H. de Jeu. New York: Gordon and Breach (1980).

7. Fundamentals of Amorphous Semiconductors edited by M. Kikuchi and K. Tanaka. Tokyo: Ohm-sha (1982) (in
Japanese).

8. Electro-Optical and Magneto-Optical Properties of Liquid Crystals by L.M. Blinov. Chichester, New York,
Brisbane, Toronto. Singapore: John Wiley & Sons Ltd. (1983).

9. Hydrogenated Amorphous Silicon: Part C, Electronic and Transport Properties edited by J.I. Pankove
(Semiconductors and Semimetals, Vol. 21 edited by R.K. Willardson and A.C. Beer). Orlando: Academic Press (1984).

10. Liquid Crystals by S. Iwayanagi. Tokyo: Kyoritsu Syuppan, Ltd. (1986) (in Japanese).

11. Hydrogenated Amorphous Silicon by R.A. Street. Cambridge: Cambridge University Press (1991).

12. Liquid Crystalline Materials edited by S. Kusabayashi. Tokyo: Kodansha, Ltd. (1991) (in Japanese).

Index

Note on Index Page Hyperlinks

This Index retains the "Print Book Page Numbers" as links to embedded targets within the content. Navigating from a "Page Number" link will
take you to within three Mobipocket Reader "Page Forward" clicks of the original Index reference point.

This strategy retains the full value of the academic Index, and presents the relative positions and distribution of Index references within this book.
The Index Page Numbers are hyperlink pointers and have no relationship to the Mobipocket Reader soft-generated page numbers.

Absorption edge, 116

Acousto-electric voltage, 115

Activation energy, 60, 68, 103, 107–108, 111–112

Additional capacitor (Cadd), 10, 13

Aluminum (Al), 55, 74, 76, 78–9

metallization, 30, 32

resistivity, 30
Alignment layer, 15

Aluminum oxide (A12 O3 ), 51, 76, 78, 80

Alphanumeric display, 51

Ammonia, 74

Amorphous silicon hydride, 120

Anisotropic dielectric constant, 139, 161

Anisotropic medium, 141, 168

Anodic oxidation, 76, 78–9

Anti-bonding states, 99

Aperture ratio, 9, 13, 26, 80

Argon, 42, 74

a-Si:H

absorption coefficient, 65

conductivity, 96, 107

deep trap levels, 69

density of states, 103

drift mobility, 113

permittivity, 83

pin diodes, 96

refractive indices, 123

specific dielectric constant, 84

a-Si:H TFT

back-channel-etched (BCE), 12, 44, 56, 74, 76

channel-passivated (CHP), 12, 56, 76

C-V characteristies, 86

high-frequency performance, 60

I-V characteristics, 66–7, 113

output characterisitics, 18, 59–60

threshold voltage (Vt), 19, 51, 59–60, 86, 91, 113


transconductance, 51, 59

transfer characteristics, 23, 51, 60, 66

Azimuthal angle, 146

Azoxy, 161

Backlight, 9, 18

illumination, 9, 13, 38, 63

spectral distribution, 36–7

Band tail, 103, 107

Bend stress, 135

Bending modes, 123

Benzene rings, 161

Biphenyl core, 161

Birefringence, 2, 141, 155, 167, 170

Black matrix, 13–5

Blue mode, 155

Bonding state, 99

Boron, 96, 111

Broken bond, 103

Buslines, 8

data, 9–10, 13, 41–2

gate, 9, 11, 13, 18, 22, 24, 26, 32, 41–2

Capacitance

additional, 13

drain-gate, 88

gate, 57–8

gate busline, 13

gate insulator, 19, 113

liquid-crystal cell, 11
parasitic, 16, 24–6, 61–2, 76, 80, 137

source-gate, 88

storage, 13, 26 188

Capacitive coupling, 14

Cathode-ray-tube (CRT), 4, 41, 137, 149

CdSe TFT, 1, 51

CF4 , 74

Channel length, 56, 57, 83, 91–2, 113

Channel width, 62, 90, 113

Chiral nematic, 154, 162

Chiral pitch, 156

Chloric acid, 42

Cholesteric benzoate, 2

Cholesteric liquid-crystals, 128

Chromaticity diagram, 32, 34–5, 38

Chromium (Cr), 14, 30, 55, 74, 79

Chromium oxide, 14

CIE, 32

Coherence length, 172

Color filter, 6, 9–10, 15, 36–8

Common electrode, 15–6, 25

Compressive stress, 69

Conduction band, 82, 103, 106, 116, 118

edge, 108, 112, 116

Contact linear sensor array, 96

Continuous random network (CRN), 102, 103

Contrast ratio, 1, 15, 130–1, 144–5, 154, 156

Coordination number, 97

Corning 7059 glass, 41


Coplanar TFT structure, 57

Copying machines, 63, 96

Covalent bonding, 97

Crystalline semiconductors, 96, 116

Crystalline silicon, 57, 60, 82, 101–2, 121

Curvature strain, 132, 172, 174–5

CVD, 74, 80

Cyanobase, 161

Cyano-pentyl-biphenyl, 161

Dangling bond, 96, 99, 102–3, 120

dc voltage offset, 16, 18, 24, 63, 137, 139

Deep-level transient spectroscopy (DLTS), 105–6

Defect density, 70

Demarcation energy, 116

Density of states, 60, 103–8, 111, 118

Depletion-type TFT, 51

Design rule, 13, 44, 56

Diagonal size, 30, 32

Diborane, 54, 111

Dielectric anisotropy, 133, 146, 167

Dielectric permittivity, 139–140, 167, 169, 170, 172

Disclination, 15, 149

Domain boundary, 149

Donor density, 83–4

Drain path, 89–91

Dry process, 76, 80

Duty ratio, 72

Dynamic range, 51, 60


Dynamic scattering, 1

Elastic constant, 167, 174, 177

Electroluminescent display, 149

Electron escape frequency, 116

Ellipticity, 141–2

ENDOR, 99

Enhancement-type TFT, 51

Energy gap, 107

ESR, 99

Euler's equation, 133, 135

Evaporated amorphous silicon, 105

EXAFS, 99

Extended state, 82, 103–4, 107, 111, 117–8

Extraordinary ray, 170

Extraordinary refractive indices, 141, 146

Fermi energy, 103–4, 107–8, 111

Flat band voltage, 84

Flat-panel display, 1, 4, 41

Flicker, 24

Fluorescent lamp, 9

Frame rate, 22, 71

Free energy, 132–3, 135, 174, 179

Freedericksz transition, 132

Gain-bandwidth product, 60

Gate

capacitance, 86

delay, 30

electrode, 24, 66, 74, 78, 80, 84, 88

metal, 76
insulator, 30, 56, 68, 72, 76, 83, 113

pulse, 20, 25, 27

Gauss's law, 83, 86

Gelatin, 37

Glow-discharge, 120, 122

Gradual channel approximation, 18, 57, 90

Gray scales, 149–150, 154

Helical axis, 162

Herring-bone-type packing, 162

HF solution, 42

Hillocks, 78–9

Hole blocking contact, 65

Homeotropic, 131–2, 141, 172

Homogeneous, 127, 131–2, 172

Hopping conduction, 107–8

Hybridized sp 3 orbitals, 99

Hydrogen, 55, 96, 101, 103, 120, 123

evolution, 120, 123

Image sticking, 24

Impurity diffusion, 58

Indirect transition, 116

Infrared absorption spectroscopy, 120

Insulated-gate field-effiect-transistors, 59

Intrinsic layer (i-layer) of a-Si:H, 56, 62, 76

Integrated circuits, 4

Inverted staggered-electrode, 56

Ion implantation, 58, 81

Ionic crystal, 97
Irrotational flow, 182

ITO (Indium Tin Oxide), 6, 13, 42

Laminar flow, 183

Laplace equation, 83

Laplace transform, 28–9

Large-scale integration (LSI), 39–40

metallization, 78

Laser-beam printer, 96

Leslie coefficients, 182, 185

Lift-off technique, 80

Light-guide, 9

Light shield, 57, 64

LiNbO 3 , 113, 115

Line-at-a-time, 10, 26

Liquid-crystal cell

capacitance, 139

resistance, 20, 56

resistivity, 23–4

transmittance, 36

threshold voltage (Vth), 24, 133, 139, 169

Lithography, 40

Localized state 83–5, 88, 103, 105, 107

Lone-pair electrons, 99

Long-range order, 97, 102, 158, 162

Luminance, 33

Lyotropic, 159

Magnetic susceptibility, 177

Mauguin's limit, 142

MBBA, 159, 161, 169, 172, 183, 185


Mesophase, 158–9, 161

Metal-Insulator-Metal (MIM), 76–7

Meyer-Neldel law, 108

Microcrystalline silicon, 45, 81

Miesowicz coefficients, 183

Minimum metallic conductivity, 104, 108

Mo, 55, 76

Mobility

a-Si:H, 2

carrier, 57, 60, 107–8, 111, 113

drift, 92, 111–3

edge, 60, 85, 103–4, 108

electron, 19, 83, 85

extended state electron, 60, 112

free carrier, 108, 111

field effect, 51

free electron, 116

free hole, 116

gap, 107, 119

hole, 60, 82

Monohydride, 123

MOS FETs, 69, 88, 113

Multiplexing, 154

n-type (n + ) a-Si:H layer, 65, 74

Nematic phase, 149, 159, 161–2, 165, 170, 176

Nitrogen, 74

Nitric acid, 42

Nitric acid cerium ammonium, 74


NMR, 99

Node potential, 18–20, 24

Non-enantiomorphic, 174–5

Non-polar, 174–5

Normally black (NB), 130, 143–4

Normally white (NW), 129, 131, 141, 143–5

NTSC, 36

Off-current, 51, 60

Ohmic contact, 57, 74, 80

On-current, 51, 60

On/off ratio, 23, 60

One-to-one projection, 40

Optical activity, 141

Optical band gap, 119

Optical lithography system, 40

Order parameter, 164, 166–7

Ordinary ray, 170

Ordinary refractive indices, 141

Orientational order, 158–9

P-azoxyanisole (PAA), 159, 161, 172, 185

Pair distribution function, 99–100

Passive-matrix LCDs, 154

Pauli exclusion principle, 97

PECVD (plasma enhanced CVD), 55

Permittivity tensor, 168

Phosphine, 54, 109

Phosphorus, 96

Photocarriers, 65, 70

Photoconductivity, 63, 76
Photocurrent, 12, 15, 65

Photoresist, 41

Photosensors, 63

Picture element (pixel), 1, 4, 8–9, 38

aperture ratio, 80

capacitance, 18, 22

electrode, 12–3, 15–6, 24, 42, 139

Piezoelectric crystal, 115

Pigment-dispersed resin, 37

Pinch-off, 59

Pixel see Picture element

Planar alignment, 127, 149

Plasma chemical vapor deposition (p-CVD), 39, 54, 74, 76, 96

Plasma-display, 149

Platinum, 79

Polarization tensor, 168

Polarizer, 140, 147–8

crossed, 7, 141, 143–4, 158

front, 2, 155

parallel, 143–4

rear, 2, 155

Polycrystalline semiconductor, 108

Polyhydride, 120

Polyimide, 7, 15, 127, 148

Poole-Frenkel.current, 76

Pretilt angle, 7, 149, 151–2, 154–6

Primary color, 32, 36

Purple boundary, 35
Quad arrangement, 38

Radial distribution function (RDF), 99, 101–2

Raman scattering, 99

Raman spectroscopy, 120, 123

Reduced cell gap, 142

Reflective mode, 1

Resolution, 4, 13, 22, 30, 44

Response time

STN cell, 156

TN cell, 136

Retention, 10, 23

Reverse tilt, 15, 149

rf

frequency, 54

glow discharge, 96

plasma, 55

power, 54–5, 68, 74, 123

RGB dots, 5

Rocking modes, 123

Rotational flow, 182

Rotational viscosity, 136, 185

Scan line, 22–23, 154

Schlieren textur, 149

Schiff base, 161

Schottky diodes, 105

Se, 112

Self-alignment process, 26, 80–1

SF6 , 74

SiH bond, 123


Shear stress, 185

Shear torque coefficient, 185

Short-range order, 97, 102, 158

Silane (SiH 4 ), 54–5, 74, 96

Silicon nitride (SiN), 55, 62, 69–70, 74–6, 79, 91

dielectric constant, 30

gate insulator, 30, 55, 69

permittivity, 91

silicon-rich, 69

SiO2 , 51

Smectic liquid-crystal, 159

Solar cells, 63, 96

Source path, 89–91

Spacer, 7

Space-charge-limited current, 89

Spectral locus, 34–5

Splay, 135, 174, 177

Stacked-layer system, 76, 79

Staebler-Wronski effect, 70

Staggered-electrode structure, 57, 82

Step-and-repeat aligners, 40

Stoichiometric composition, 69

Storage capacitor (Cst), 1, 10–1, 15

Stress tensor, 181–2

Stretching-modes, 123

Strong anchoring, 132, 135, 149, 172

Subpixel, 5

Substrate,
color filter, 6, 13–5, 18

common electrode, 6–7

glass, 2, 5–7, 13, 15, 37, 39, 139, 148, 172

TFT, 5, 7–8

Subthreshold slope, 60

Surface acoustic wave (SAW), 111, 113, 115–6

Supertwisted birefringent effect (SBE), 154

Susceptibility tensor, 167–8

Ta, 55, 76

Ta2 O5 , 76–7

TAB (Tape Automated Bonding), 8

Tail state, 60, 103, 111

Tauc edge, 119

Tensile stress, 69

Texture patterns, 159

Thermal CVD systems, 55

Thermal release rate, 115

Thermotropic, 159

Three-teminal devices, 53

Three-wave-length-type illumination, 9, 37

Threshold voltage (Vth ), 139, 169

STN display, 155–6

TN cell, 131, 133, 135–6

Threshold voltage shift (Vt shift, /_\V t), 67–72

Throughput, 41

Tilt angle, 131, 134, 146

Time-of-flight measurement, 112, 116

Total stimulus, 33

Transistor, 4
Transit time, 113

Transmissive mode, 1

Transparent electrode, 18

Trapped electron density, 82–3

Triad arrangement, 38

Tristimulus values, 32

Turbulent liquid, 2

Twist, 135, 174, 177

Twist angle, 142, 154, 156

Two-domain TN cell, 150

Two-terminal devices, 53

Uniaxial crystal, 170

Urbach edge, 120

Vacuum tube, 89

Valence band, 99, 103, 106, 111, 116, 118

Velocity-gradient tensor, 181–182

Vickers hardness, 55

Video-graphic-array (VGA), 4, 13, 71

Viewing-angle characteristics, 146, 148, 150, 152

Viscosity coefficient, 167, 181, 183, 185

Viscous-stress tensor, 181, 183

Vitreous semiconductors, 103

Wagging modes, 123

Whiskers, 78

X-ray diffraction, 99, 101

Yellow mode, 155

Yield, 13

You might also like