Amorphous Semiconductors - Kugler and Shimakawa
Amorphous Semiconductors - Kugler and Shimakawa
Amorphous Semiconductors - Kugler and Shimakawa
S Á N D O R K U G L E R
Budapest University of Technology and Economics, Hungary
KO I C H I S H I M A K AWA
University of Pardubice, Czech Republic
University Printing House, Cambridge CB2 8BS, United Kingdom
www.cambridge.org
Information on this title: www.cambridge.org/9781107019348
c S. Kugler and K. Shimakawa 2015
This publication is in copyright. Subject to statutory exception
and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without the written
permission of Cambridge University Press.
First published 2015
Printed and bound in the United Kingdom by CPI Group Ltd, Croydon CR0 4YY
A catalog record for this publication is available from the British Library
Library of Congress Cataloging in Publication data
Kugler, Sándor, 1950–
Amorphous semiconductors / Sándor Kugler, Budapest University of Technology and Economics,
Hungary, Koichi Shimakawa, University of Pardubice, Czech Republic.
pages cm
Includes bibliographical references and subject index.
ISBN 978-1-107-01934-8 (hardback)
1. Amorphous semiconductors. 2. Semiconductors – Computer simulation.
I. Shimakawa, Koichi. II. Title.
QC611.8.A5K84 2015
537.6 223 – dc23 2013022108
ISBN 978-1-107-01934-8 Hardback
Cambridge University Press has no responsibility for the persistence or accuracy of
URLs for external or third-party internet websites referred to in this publication,
and does not guarantee that any content on such websites is, or will remain,
accurate or appropriate.
To my wife, Szabina, and our children, Szilvia and Zsófia
SK
Preface page ix
1 Introduction 1
1.1 Historical overview: science and applications 1
1.2 Definitions 6
References 10
2 Preparation techniques 14
2.1 Growth of thin-film forms 14
2.2 Melt-quenched glasses 18
2.3 Other techniques 19
2.4 The Phillips constraint theory 20
References 22
3 Structure 24
3.1 Differences between amorphous and crystalline
semiconductors 24
3.2 Projection from three-dimensional structures to
one-dimensional functions 33
3.3 Three-dimensional structure derived from one-dimensional
function 41
3.4 Phase change and its applications 73
References 78
4 Electronic structure 86
4.1 Chemical bonds 86
4.2 Electronic density of states 90
viii Contents
4.3 Defects 93
4.4 Optical and electronic properties 104
References 113
Index 145
Preface
1
2 Introduction
were later applied to laser printers. More recently, a-Se has been replaced by
organic polymers for xerography (Weiss and Abkowitz, 2006). The main reason
for this is that organic materials cost less.
Most people would recognize the acronym “DVD” (digital versatile disk).
The basic operation of the DVD was first proposed by Feinleib et al. (1971),
although the material Ge2 Sb2 Te5 (also known as GST) has been employed
commercially by the Panasonic group in Japan (Yamada et al., 1991). A DVD
operates via optically induced phase changes (and hence changes in the reflection
coefficient) between amorphous and crystalline states. Using GST, DVDs can be
rewritten in excess of one million times, with a crystallization time of less than
50 ns achieved during each rewriting process. The DVD system currently has a
memory capacity exceeding 50 GB per disk using a blue semiconductor laser.
In the near future, rewritable electrical memory devices will be commercially
available that use a phase-change random access memory (PRAM), following the
memory switching devices proposed by Ovshinsky (1968). In addition, phase-
change materials offer a promising route for the practical realization of new
forms of general-purpose and “brain-like” computers that could learn, adapt,
and change over time (Wright et al., 2011).
As a-Se is a very sensitive photoconductor, especially for x-rays, due to its
high atomic weight, it has been possible to realize a direct x-ray imaging device
for use in the medical field. The image of a human hand by Rowlands and Kasap
(1997) recalls Wilhelm Röntgen’s first x-ray photograph of his wife’s hand. This
device incorporates a large area of thick (1 mm) a-Se evaporated onto a thin-film
transistor (TFT) made from a hydrogenated amorphous silicon (a-Si:H) active
matrix array (AMA). The x-ray-induced carriers in a-Se travel along the electric
field lines and are collected at their respective biased electrode and storage
capacitor. The stored images are then sent directly to the medical specialist’s
computer. This type of x-ray image sensor is used widely in mammography.
The Japan Broadcasting Corporation, NHK, has utilized the “avalanche
photomultiplication” effect in a-Se (Juška, Arlauskas, and Montrimas, 1987;
Tanioka, 2007) to create a powerful broadcasting tool. High-gain avalanche
rushing amorphous photoconductor (HARP) vidicon tubes have been developed
by K. Tanioka and his collaborators, leading to the construction of a HARP
vidicon TV camera that is over 100 times more sensitive than a CCD camera.
(2) Hydrogenated amorphous silicon (a-Si:H). In the early 1970s, the oil
crisis in the Middle East led to the serious consideration of using photovoltaic
(PV) cells as an alternative source of energy. To use PV cells as a viable power
1.1 Historical overview 3
source requires a large area and low cost. Consequently, the first a-Si:H solar
cells were fabricated by Carlson and Wronski (1976) at RCA Laboratories in
Princeton, NJ. Later, Y. Kuwano’s group at Sanyo Co. Ltd. (Japan) was the first to
market the PV devices. Common structures comprise p-i-n type heterojunctions.
There are several types of PV cells available, the most common being tandem
(with dual and triple junctions) a-Si:H configurations, for which more than 10%
efficiency is achieved in large-area commercial devices (Carlson et al., 1996).
People have recognized the importance of developing PV devices following
the Fukushima nuclear power station disaster caused by the earthquake and
subsequent tsunami on 11 March 2011.
Thin-film transistors (TFTs) using a-Si:H were first developed in the form
of field effect transistors (Powell, 1984; Spear and LeComber, 1984). Two of
the most important requirements for TFTs are a high ON/OFF current ratio and
a small gate voltage; these are achieved by using a-Si:H. These characteristics
mean that TFTs are suitable for use as switching transistors in a liquid crystal
display (TFT-LCD), which has completely replaced the former cathode ray
tube. Following subsequent improvements in TFT-LCDs, flat-panel displays
(FPDs) now produce the clear and large (over 100 cm) images used in TVs and
monitors.
(3) Oxides. The high demand for flexible and optically transparent TFTs for
use in the next-generation FPDs led to the realization of transparent conductive
oxides (TCOs) by H. Hosono’s group in 1996 (see, for example, Hosono (2006)).
Transparent TFTs were developed using ionic oxides such as a-InGaZnO4
(known as a-IGZO). The electron mobility of a-IZGO is larger than that of
a-Si:H, and the TFT stability is excellent. Samsung’s group in Korea developed
the a-IGZO TFT-LCD display that is used commercially in the iPad 3 (Apple
Inc.). The high quality and stability of a-IGZO TFT-LCD large-area displays
may dominate the “display world” in the near future.
It should be noted that two great discoveries – the amorphous chalcogenides
(known as a-chalcogenides or a-Chs) used in electrical memory (Ovshinsky,
1968) and switching devices, and device-quality hydrogenated amorphous sil-
icon (a-Si:H) (Spear and LeComber, 1975) – initiated a vast field of science,
which will be briefly reviewed in the following.
The most important scientific issue that resulted from work on amorphous
semiconductors was the structural information gained from experiment and the-
ory (in the form of modeling), because the fundamental physical and chemical
properties are principally determined by their structure. For tetrahedrally bonded
4 Introduction
is dominated by the DOS. Street (1976) initiated work on PL in this field, and
recently the technique of PL with wide-lifetime distribution (in the nanosecond
to millisecond range) has become well established (Aoki, 2012).
Thermally activated band-type electronic transport occurs in both a-Si:H and
amorphous chalcogenides near room temperature. The Meyer–Neldel compen-
sation law for thermally activated processes has been discussed with reference to
disordered matter; however, this compensation law is not clearly understood from
a theoretical standpoint (Yelon, Movaghar, and Crandall, 2006). At low tempera-
tures, or in defect-dominated materials, transport through localized states is
dominant. Mott (1969) developed the variable-range-hopping (VRH) model for
a single-phonon carrier hopping process. The importance of small polarons
has been pointed out in disordered materials (Emin, 1975), in which multi-
phonon processes dominate transport. Whether multiphonon or single-phonon
processes dominate in amorphous semiconductors is still a matter of debate
(Shimakawa and Miyake, 1988; Emin, 2008). Recent results in hydrogenated
amorphous/nanocrystalline Si suggest that this factor will depend on tempera-
ture (Wienkes, Blackwell, and Kakalios, 2012). Alternating current (ac) transport
studies, initiated by Pollak and Gegalle (1961) in crystalline Si, provide infor-
mation on localized states. Theories involving the continuous time random walk
(CTRW) of carriers have developed well, and apply to hopping systems (Dyre
and Schröder, 2000).
Time-of-flight measurements developed by Spear and Adams (1966) were
applied to crystalline sulfur. In amorphous semiconductors, this technology has
been applied to many systems involving a-Si:H and a-Chs to study the drift
mobility of both electrons and holes. In most cases, non-Gaussian transport
has been discovered. This opened the door for studying dispersive transport in
disordered matter (Pfister and Scher, 1978).
The illumination of amorphous semiconductors, effectively with bandgap
light, induces various changes on structural and electronic properties. In amor-
phous chalcogenides, the bandgap decreases with illumination; this is referred to
as photodarkening, which was first reported by DeNeufville, Moss, and Ovshin-
sky (1973/1974), and it was confirmed that photodarkening accompanied volume
changes (Hamanaka et al., 1976). These effects are of interest in scientific appli-
cations, and hence a huge amount of work has been devoted to these topics
(see the reviews by Tanaka (1990), Pfeiffer, Paesler, and Agarwal (1991), and
Shimakawa, Kolobov, and Elliott (1995)). Photoinduced volume changes in a-Se
films have been discussed using MD simulations (Hegedüs et al., 2005). Giant
6 Introduction
1.2 Definitions
Crystalline and non-crystalline structure
For the reader unfamiliar with amorphous semiconductors, we present here a
brief explanation of the most essential definitions. Firstly, we must stress that
there is a considerable amount of confusion in the scientific literature concerning
the terms non-crystalline, amorphous, glassy, vitreous, randomness, disorder,
liquid, and even crystalline. A first important question is whether an atomic
structure is crystalline or non-crystalline (amorphous). A perfect crystal is one
in which the atoms, or a group of atoms, are arranged periodically in three
dimensions to an infinite extent and are rigidly fixed at their thermal equilibrium.
This mathematical model of atomic configurations provides us with a relatively
easy method for calculating the different properties of condensed matter, which
may be found in books on theoretical solid state physics.
A more realistic arrangement is an imperfect crystal in which the atoms
form a pattern that repeats periodically only to a finite extent. In fact, this may
seem more realistic, but we still have a serious problem in that this type of
imperfect crystal behaves counter to thermodynamics. This is because, at non-
zero temperatures, defects form in any atomic configuration when the crystals
are in their equilibrium states. Real crystals are, of course, not only finite in
size, but also contain imperfections such as vacancies, interstitial (foreign or
self) atoms, dislocations, impurities, distortions associated with the surface, etc.
Furthermore, another effect must be taken into account at finite temperature. The
random motion of atoms at their equilibrium positions also weakly destroys the
perfect periodicity at any given moment in time. These defects cause distortions
1.2 Definitions 7
in the crystal lattice, but we do not consider these real crystals to be amorphous
solids. Translational symmetry more or less remains.
Until 1992, a crystal was defined by the International Union of Crystallog-
raphy as “a substance, in which the constituent atoms, molecules, or ions are
packed in a regularly ordered, repeating three-dimensional pattern.” In 1984 (two
years after their discovery), Shechtman et al. (1984) published a paper on rapidly
solidified alloys of aluminum with 10–14% manganese that possess icosahedral
symmetry in combination with long-range order, named quasicrystals, in clear
violation of the above definition of a crystal. Since this discovery, hundreds of
similar atomic structures have been synthesized in laboratories around the world.
Recently, the naturally occurring quasicrystalline mineral icosahedrite has been
identified in a sample from the Khatyrka river in Chukotskii Autonomous Okrug,
Far-Eastern Region, Russia. The International Union of Crystallography has had
to modify their declaration, and the new and broader definition of crystal became
“any solid having an essentially discrete diffraction diagram.”
Other types of deviation from perfection occur in crystals. Consider a perfect
lattice, but one in which each atomic site possesses a randomly oriented spin
or other magnetic moment. Another partially ordered arrangement is possible
in materials built up of large near-spherical molecules. Molecular centers of
mass can be found in crystalline sites where the directions of molecular axes
are randomly distributed. A typical example of the so-called plastic crystals is
the solid crystalline phase of C60 molecules. A similar example occurs when
a binary (ternary) alloy has two (three) different, randomly distributed atoms
inside a perfect lattice. Are they crystalline or non-crystalline? This issue does
not occur for amorphous materials.
A liquid crystal is a state of matter that has properties between those of
a conventional liquid and those of a solid crystal. Rigid rod-shaped or disc-
like molecules develop in a well-defined direction. The centers of mass of the
molecules are distributed randomly in two or three dimensions. How does one
distinguish between the condensed phase and the liquid phase? A solid is a phase
whose shear viscosity exceeds 1013.6 Pa s (N s m−2 ), although this threshold is
rather arbitrary. Values for the viscosity of common liquids at room temperature
are on the order of 10–3 Pa s. For some examples, see Table 1.1. Liquid crystals
are anisotropic fluids.
A granular material is a collection of distinct macroscopic particles with
randomly distributed centers of mass, such as sand or peanuts in a container.
There is a weak interaction between macroscopic particles. Their behavior is
8 Introduction
differently from that of solids, liquids, and gases, and this has led many to
characterize granular materials as a new form of matter. A similar topology
can be observed in the “microworld.” Polycrystalline materials (poly is a Greek
word meaning “many”) consist solely of crystalline grains, separated by grain
boundaries, and the interaction between them is strong. Furthermore, in some
cases crystalline grains do not touch each other; rather, they are embedded
into a disordered atomic environment. The grain size determines whether these
materials are named nanocrystalline or microcrystalline materials.
Amorphous materials exist in the condensed phase and do not possess the long-
range translational order (periodicity) of atomic structures. They do not exhibit a
discrete diffraction pattern.
Using this terminology, the set of amorphous materials has within it a funda-
mental subset glasses, where a glass is an amorphous solid that exhibits a glass
transition. If a liquid (melt) is cooled very rapidly so that crystallization can
be bypassed, the disordered structure can be frozen-in. This disordered con-
densed phase is known as a glass. Such a glass-forming process involves the
1.2 Definitions 9
where ρ 0 and ε 0 are constants, kB is the Boltzmann factor, and T is the tempera-
ture. Note that ρ decreases with increasing T, which is opposite to the case for
conventional metals (where ρ is proportional to T). Most chalcogenide glasses
and hydrogenated a-Si follow the above equation, whereas pure a-Si and a-Ge
do not.
Atomic structure is one of the main features that distinguish between different
electronic transport properties. Carbon provides us with a simple example. The
diamond crystal arrangement of carbon atoms creates a good insulator, with
about a 5.5 eV gap in the electronic DOS. However, the non-crystalline config-
uration of carbon displays semiconductor properties, and the graphitic atomic
structure is a conductor. Furthermore, superconductivity at 30 K in caesium-
doped C60 has been observed.
Pure amorphous semiconductor materials are located in the even-numbered
groups of the periodic table, such as group IV (carbon, silicon, and germanium)
10 Introduction
and group VI (selenium and tellurium). In group VI, stable allotropes of sulfur
are excellent electrical insulators.
The basis of this structure was formulated by Sir Nevill Francis Mott (Mott,
1969), who concluded via his famous (8 – N) rule that there are no dangling bonds
inside such a material. This is explained as follows. In a glass, any atom appears
in such a way that it retains its natural coordination. The number of covalent
bonds Z = 8 – N, where N is the number of valence electrons. (We consider
elements only in groups IV, V, and VI.) Further, an additional rule states that
Z = N if N < 4. The most important consequence of this was that amorphous
semiconductors cannot be doped. Today, however, we know this is not the
case.
References
Aoki, T. (2012). Photoluminescence spectroscopy. In Characterization of Materials,
2nd edn., ed. E. N. Kaufmann. Hoboken, NJ: John Wiley and Sons, pp. 1158–
1169.
Carlson, D.E. and Wronski, C.R. (1976). Amorphous silicon solar cell. Appl. Phys. Lett.,
28, 671–673.
Carlson, D.E., Arya, R.R., Bennett, M. et al. (1996). Commercialization of multijunction
amorphous silicon modules, Conf. Record of the Twenty Fifth IEEE, Photovoltaic
Specialists Conf., Crystal City, Washington D.C., May 13–17, 1996, pp. 1023–1028.
Cohen, M.H., Fritzshe, H., and Ovshinsky, S.R. (1969). Simple band model for amor-
phous semiconducting alloys. Phys. Rev. Lett., 22, 1065–1068.
DeNeufville, J.P., Moss, S.C., and Ovshinsky, S.R. (1973/1974). Photostructural trans-
formations in amorphous As2 Se3 and As2 S3 films. J. Non-Cryst. Solids, 13, 191–
223.
Drabold, D.A., Zhang, X., and Li, J. (2003). First-principles molecular dynamics and
photostructural response in amorphous silicon and chalcogenide glasses. In Photo-
Induced Metastability in Amorphous Semiconductors, ed. A.V. Kolobov. Weinheim:
Wiley-VCH, pp. 260–278.
Dyre, J.C. and Schröder, T.B. (2000). Universality of AC conduction in disordered
solids. Rev. Mod. Phys., 72, 873–892.
Emin, D. (1975). Phonon-assisted transition rate: optical-phonon-assisted hopping in
solids. Adv. Phys., 24, 135–218.
Emin, D. (2008). Generalized adiabatic polaron hopping: Meyer-Neldel compensation
and Poole-Frenkel behaviour. Phys. Rev. Lett., 100, 16602–16604.
Feinleib, J., deNeufville, J., Moss, S.C., and Ovshinsky, S.R. (1971). Rapid reversible
light-induced crystallization of amorphous semiconductors. Appl. Phys. Lett., 18,
254–257.
Greaves, G.N. and Davis, E.A. (1974). A continuous random network model with three-
fold coordination. Philos. Mag., 29, 1201–1206.
Greaves, G.N. and Sen, S. (2007). On organic glasses, glass-forming liquids and amor-
phizing solids. Adv. Phys., 56, 1–166.
References 11
Hamanaka, H., Tanaka, K., Matsuda, A., and Iizima, S. (1976). Reversible photo-induced
volume changes in evaporated As2 S3 and As4 Se5 Ge1 films. Solid State Commun.,
19, 499–501.
Hegedüs, J., Kohary, K., Pettifor, D.G., Shimakawa, K., and Kugler, S. (2005). Photo-
induced volume changes in amorphous selenium. Phys. Rev. Lett., 95, 206803, 1–
4.
Hisakuni, H. and Tanaka, K. (1995). Optical microfabrication of chalcogenide glasses.
Science, 270, 217–218.
Hosono, H. (2006). Ionic amorphous oxide semiconductors: material design, carrier
transport, and device application. J. Non-Cryst. Solids, 352, 851–858.
Juška, G., Arlauskas, K., and Montrimas, E. (1987). Features of carriers at very high
electric fields in a-Se and a-Si:H. J. Non-Cryst. Solids, 97–98, 559–561.
Kastner, M.A. (1972). Bonding bands, lone-pair bands, and impurity states in chalco-
genide semiconductors. Phys. Rev. Lett., 28, 355–357.
Kastner, M.A., Adler, D., and Fritzshe, H. (1976). Valence-alternation model for local-
ized gap states in lone-pair semiconductors. Phys. Rev. Lett., 37, 1504–1507.
Long, M., Galison, P., Alben, R., and Connell, G.A.N. (1976). Model for structure of
amorphous selenium and tellurium. Phys. Rev. B, 13, 1821–1829.
Morigaki, K. (1999). Physics of Amorphous Semiconductors. London: Imperial College
Press and World Scientific.
Mott, N.F. (1969). Conduction in non-crystalline materials. Phil. Mag., 19, 835–852.
Mott, N.F. (1992). Conduction in Non-Crystalline Materials, 2nd edn. Oxford: Claren-
don Press.
Mott, N.F. and Davis, E.A. (1979). Electronic Processes in Non-Crystalline Materials,
2nd edn. Oxford: Clarendon Press.
Ovshinsky, S.R. (1968). Reversible electrical switching phenomena in disordered struc-
tures. Phys. Rev. Lett., 21, 1450–1453.
Pai, D.M. and Springett, B.E. (1993). Physics of electrophotography. Rev. Mod. Phys.,
65, 163–211.
Pfeiffer, G., Paesler, M.A., and Agarwal, S.C. (1991). Reversible photodarkening of
amorphous arsenic chalcogens. J. Non-Cryst. Solids, 130, 111–143.
Pfister, G. and Scher, H. (1978). Dispersive (non-Gaussian) transient transport in dis-
ordered solids. Adv. Phys., 27, 747–798.
Phillips, J.C. (1979). Topology of covalent non-crystalline solids I: short-range order in
chalcogenide alloys, J. Non-Cryst. Solids, 34, 153–181.
Polk, D.E. (1971). Structural model for amorphous silicon and germanium. J. Non-Cryst.
Solids, 5, 365–376.
Pollak, M. and Gegalle, T.H. (1961). Low-frequency conductivity due to hopping pro-
cesses in silicon. Phys. Rev., 122, 1742–1753.
Powell, M.J. (1984). Material properties controlling the performance of amorphous
silicon thin film transistors. MRS Symp. Proc., 33, 259–274.
Rowlands, J.A. and Kasap, S.O. (1997). Amorphous semiconductors usher in digital
X-ray imaging. Phys. Today, 50, 24–31.
Redfield, D. and Bube, R.H. (1996). Photoinduced Defects in Semiconductors.
Cambridge: Cambridge University Press.
Selényi, P. (1935a). Elektrographie, ein neues elektrostatisches Aufzeichnungsverfahren
und seine Anwendungen. Elektrotech. Z., 56(35), 961–963.
Selényi, P. (1935b). Methoden, Ergebnisse und Aussichten des elektrostatischen Auf-
zeichnungsverfahrens. (Elektrographie). Z. Tech. Phys., 12, 607–614.
12 Introduction
Yamada, N., Ohno, E., Nishiuchi, K., Akahira, N., and Takao, M. (1991). Rapid phase-
transitions of GeTe-Sb2 Te3 pseudobinary amorphous thin films for an optical disk
memory. J. Appl. Phys., 69, 2849–2856.
Yelon, A., Movaghar, B., and Crandall, R.S. (2006). Multi-excitation entropy: its role in
thermodynamics and kinetics. Rep. Prog. Phys., 69, 1145–1194.
2
Preparation techniques
14
2.1 Growth of thin-film forms 15
Thermal evaporation
This is the most conventional method used in preparing amorphous materials.
The starting materials comprise ingots or powders, which are mounted in a
sample boat and melted in a vacuum chamber at around 10–6 Torr (10–4 Pa).
To achieve this low pressure, an oil diffusion pump is used in conjunction with a
liquid-nitrogen-cooled trap. Note that the boat itself acts as the electrical heater.
High-melting-point metal boats, made for example from molybdenum (Mo) or
tungsten (W), are normally used. This technique is suitable for low-melting-
point materials, and the vapor is condensed at arbitrary temperatures onto the
desired substrates. The deposition rate, which is usually about 10 nm s−1 , is
controllable and depends on the boat temperature. The thickness of the film can
be controlled by the deposition time. Well-known amorphous chalcogenides, for
example Se, As2 Se(S)3 , GeSe(S)2 , etc., can be prepared using this technique. The
main disadvantage of thermal evaporation is in ensuring that the composition of
the resulting films is the same as that of the starting materials.
Device-quality a-Se thick (>1 mm!) films, which are commercially utilized
for direct x-ray image detectors (Rowlands and Kasap, 1997) and high-gain
avalanche rushing amorphous photoconductor (HARP) vidicon TV cameras
(Tanioka, 2007), are prepared by the traditional thermal evaporation technique.
Sputtering
The usual sputtering process comprises a radio-frequency (rf, 13 MHz) high-
electric field applied between the target (source material) and the substrate
electrodes. A sputtering gas (0.13–2.7 Pa) (for example, Ar) is introduced into a
chamber and is pumped away by a vacuum system. The substrate temperature is
easy to control. Positively ionized gases produced by the rf field supply kinetic
energy to the atoms on the surface of the source material target. Atoms or
molecules dissociated from the target are deposited onto the substrate in the form
of films. Reactive gases such as H2 and/or N2 are also used for sputtering, in a
technique known as reactive sputtering, in which the gases are incorporated into
the deposited films (for example, a-Si:H, a-C:H, a-Si1–x Nx , etc.). The deposition
rate is lower than that for evaporation techniques, i.e. less than 1 nm s−1, although
this will depend greatly on the sputtering conditions.
The sputtering process is more complicated than thermal evaporation; for
example, we must pay attention to the sputtering gas pressure, the rf power
applied to the target, the bias voltage of the target or substrate, etc. However,
sputtering is superior to evaporation for multicomponent systems, and the com-
position of the films is almost the same as that of the source materials (targets)
because the sputtering rates for different elements are on the same order.
When a magnetic field is applied perpendicular to the electrodes, deposition is
enhanced; this is because the ionized gases are confined in a spiral motion by the
magnetic field (Lorentz force). This method is called magnetron sputtering, and
it significantly improves the deposition rate. Magnetron sputtering is therefore
suitable and cost-effective for large-scale production, and is used for optically
transparent conductive oxides (TCOs) such as indium tin oxide (ITO) and semi-
conducting IGZO for the flat-panel display industry (Kamiya, Nomura, and
Hosono, 2009). Note that device-quality a-IZGO films are also prepared using
pulsed laser deposition, which will be described later. So called phase-change
materials, such as the chalcogenide Ge2 Sb2 Te5 used in optical memory devices
(such as DVDs), are also prepared using the magnetron sputtering technique.
(PECVD) (Schiff, Hegedus, and Deng, 2010). The plasma is introduced via the
application of a rf field (usually 13.6 MHz), and is therefore referred to as rf
PECVD. To prepare n or p type Si:H films, PH3 or B2 H6 gas is introduced onto
SiH4 . Radio-frequency PECVD is a very popular technique and provides high-
quality uniform films, although the deposition rate can be as low as 0.3 nm s−1 .
The development of this technique, by controlling the gas pressure (0.05–2 Torr),
rf power (10–100 mW cm−2 ), and substrate temperature (150–350 °C) etc.,
has led to growing numbers of commercial applications, such as large-area
photovoltaics and thin-film transistors.
Radio-frequency PECVD is also widely used in the preparation of hydro-
genated amorphous carbon (a-C:H) thin films and nanoparticles. Hydrocarbons,
such as CH4 or C6 H6 , are the main precursor materials. Changing the pressure
and the rf power allows the preparation of a-C films with remarkably different
characters, ranging from soft and porous polymer-like a-C:H and hard diamond-
like carbon layers to highly conductive graphitic structures (Veres, Tóth, and
Koós, 2008). At low ion energies, intact molecules of the feed gas were found
to be incorporated into the matrix of the amorphous films (Veres, Koós, and
Pócsik, 2002). Hydrogenated amorphous carbon nanoparticles are prepared by
using high gas pressures together with low ion energies and a benzene pre-
cursor (so-called “dusty plasma”) (Pócsik et al., 2003). Specific dusty-plasma
conditions allow amorphous nanoparticles and thin films to be obtained simul-
taneously (Veres et al., 2005).
When PECVD is carried out at very high frequencies (40–100 MHz) it is
called VHF PECVD; this technique improves the deposition rate, which can
reach 2 nm s−1 . VHF PECVD also produces high-quality films, although poor
uniformity is a problem. Microwave PECVD performed at 2.45 GHz (MW
PECVD) significantly enhances the deposition rate (up to 10 nm s−1 ), although
device-quality a-Si:H films are not easy to obtain. Note, however, that VHF
PECVD and MW PECVD are very useful in the preparation of microcrystalline
Si films (µc-Si:H), as the deposition rate of VHF PECVD for µc-Si:H is very
small. A high deposition rate for low-cost mass production can be a necessary
condition.
set-up for a HWCVD system is similar to that for rf PECVD, except that the rf
electrode is replaced by a heated filament. The gas introduced into the chamber
is catalytically excited, or decomposed into radicals or ions, by a metal (Pt, W,
Ta, etc.) filament heated to around 1800–2000 °C. Then, for example, Si radicals
diffuse inside the chamber and are deposited onto a substrate that is maintained
at a relatively high temperature (150–450 °C). Although device-quality films
with high deposition rates are obtained using this technique, a drawback is the
poor uniformity of deposited films.
bond direction is derived by the addition of two bond angles. Therefore the bond-
bending number is 2Z – 3. The total number of linearly independent constraints
Nc = Z2 + (2Z – 3). By the Phillips theory, Nc must be equal to the degrees of
freedom. In the three-dimensional (3D) case, Nc = Nd = 3, so
The SiO system is a well-known exception to the Phillips theory, with SiO2
instead of SiO4 being the optimum glass composition. This is because the Si–
O–Si bond-angle distribution is rather wide (Kuo, Lee, and Hwang, 2008), and
therefore the constraint associated with oxygen bond angles is weak and may
be neglected. Consider a Six O(1–x) binary alloy. In the equation 3 = Z2 + (2Z –
3), the (2Z – 3) term associated with the bond angles must be modified as there
22 Preparation techniques
is no zero-bond-angle constraint for oxygen. The revised term for bond bending
is therefore
(2ZSi −3) x + 0 (1 − x) = 5x
3 = Z/2 + 5x,
References
Borisova, Z.U. (1981). Glassy Semiconductors. New York: Plenum.
Delahoy, A.E. and Guo, S. (2010). Transparent conducting oxides for photovoltaics. In
Handbook of Photovoltaic Science and Engineering, 2nd edn., eds. A. Luque and
S. Hegedus. Chichester: John Wiley and Sons, pp. 716–796.
Elliott, S.R. (1990). Physics of Amorphous Materials, 2nd edn. Harlow: Longman Sci-
ence & Technical.
Fukunaga, T., Utsumi, M., Akatsuka, H., Misawa, M., and Mizutani, U. (1996). Structure
of amorphous Se prepared by milling. J. Non-Cryst. Solids, 205–207, 531–535.
Impellizzeri, G., Mirabella, S., and Grimaldi, M.G. (2011). Ion implantation damage
and crystalline-amorphous transition in Ge. Appl. Phys. A, 103, 323–328.
Kamiya, T., Nomura, K., and Hosono, H. (2009). Origins of high mobility and low
operation voltage of amorphous oxide TFTs: electronic structure, electron transport,
defects and doping. J. Disp. Technol., 5, 468–482.
Koch, C.C., Calvin, O.B., Macklamey, C.G., and Scarbrough, J.O. (1983). Preparation of
‘‘amorphous’’ Ni60 Nb40 by mechanical alloying. Appl. Phys. Lett., 43, 1017–1019.
Koffel, S., Claverie, A., BenAssayag, G., and Scheibin, P. (2006). Amorphization kinetics
of germanium under ion implantation. Mater. Sci. Semicond. Process., 9, 664–667.
Kuo, C-L., Lee, S., and Hwang, G.S. (2008). Strain-induced formation of surface defects
in amorphous silica: a theoretical prediction. Phys. Rev. Lett., 100, 076104, 1–4.
Laaziri, K., Kycia, S., Roorda, S. et al. (1999). High resolution radial distribution
function of pure amorphous silicon. Phys. Rev. Lett., 82, 3460–3463.
Mahan, A., Xu, Y., Nelson, B. et al. (2001). Saturated defect densities of hydrogenated
amorphous silicon grown by hot-wire chemical vapour deposition at rates up to
150 Å. Appl. Phys. Lett., 78, 3788–3790.
Mott, N.F. (1969) Conduction in non-crystalline materials. Phil. Mag., 19, 835–852.
Mott, N.F. and Davis, E.A. (1979). Electronic Processes in Non-Crystalline Materials,
2nd edn. Oxford: Clarendon Press.
Phillips, J.C. (1979). Topology of covalent non-crystalline solids I: short–range order in
chalcogenide alloys. J. Non-Cryst. Solids, 34, 153–181.
Pócsik, I., Veres, M., Füle, M. et al. (2003). Carbon nano-particles prepared by ion-
clustering in plasma. Vacuum, 71, 171–176.
Popescu, M.A. (2000). Non-Crystalline Chalcogenides. Dordrecht: Kluwer.
References 23
Ridgway, M.C., Azevedo, G. de M., Glover, C.J., Yu, K.M., and Foran, G.J. (2003).
Common structure in amorphised compound semiconductors. Nucl. Inst. Meth.
Phys. Res. B, 199, 235–239.
Rowlands, J.A. and Kasap, S.O. (1997). Amorphous semiconductors usher in digital
X-ray imaging. Phys. Today, 50, 24–31.
Schiff, E.A., Hegedus, S., and Deng, X. (2010). Amorphous silicon-based solar cells. In
Handbook of Photovoltaic Science and Engineering, 2nd edn., eds. A. Luque and
S. Hegedus. Chichester: John Wiley and Sons, pp. 487–545.
Schwarz, R.B. and Koch, C.C. (1986). Formation of amorphous alloys by the mechanical
alloying of crystalline powders of pure metals and powders of intermetallics. Appl.
Phys. Lett., 49 146–148.
Shimakawa, K., Hayashi, K., Kameyama, T., and Watanabe, T. (1991). Anomalous
electrical conduction in graphite-vaporized films. Phil. Mag. Lett., 64, 375–378.
Spear, W.E. and LeComber, P.G. (1975). Substitutional doping of amorphous silicon.
Solid State Commun., 17, 1193–1196.
Tanaka, K. (1989). Structure phase transitions in chalcogenide glasses. Phys. Rev. B, 39,
1270–1279.
Tanaka, K. and Shimakawa, K. (2011). Amorphous Chalcogenide Semiconductors and
Related Materials. New York: Springer.
Tanioka, K. (2007). The ultra sensitive TV pickup tube from conception to recent
development. J. Mater. Sci., 18, S321–S325.
Veres, M., Füle, M., Tóth, S. et al. (2005). Simultaneous preparation of amorphous solid
carbon films, and their cluster building blocks. J. Non-Cryst. Solids, 351, 981–986.
Veres, M., Koós, M., and Pócsik, I. (2002). IR study of the formation process of hydro-
genated amorphous carbon film. Diam. Relat. Mater., 11, 1110–1114.
Veres, M., Tóth, S., and Koós, M. (2008). New aspects of Raman scattering in carbon-
based amorphous materials. Diam. Relat. Mater., 17, 1692–1696.
3
Structure
24
3.1 Classification of semiconductors 25
and Johnson, 1982; Kramer, King, and Mackinnon, 1983; Brey, Tejedor, and
Verges, 1984; Winer and Cardona, 1986; Kugler and Náray-Szabó, 1987; Bose,
Winer, and Andersen, 1988; Kugler, Surjan, and Náray-Szabó, 1988; Kugler and
Náray-Szabó, 1991a, 1991b).
Net atomic charges are not observable directly, but experimental determin-
ation of their fluctuation is possible. By considering the deviation from the
ideal bond angle, a simple model was developed for the derivation of net
atomic charges in a-Si and diamond-like a-C (da-C) (Kugler et al., 1988;
Kugler and Náray-Szabó, 1991b). Consider an elementary triad of atoms,
denoted by K, L, and M, forming two bonds (KM and LM), with the angle
KML denoted by θ (see Figure 3.1). The net charge on the atoms of the
triad depends linearly on the deviation of θ from the ideal tetrahedral value
(dθ = θ – 109.47°):
3 3
3
3
3 2 L K 2
3
θ
1 M
x θi
y
3 2 θj 2 3
3
3 3
3
z
ΔQ
40
30
20
10
兺
–50 50
–10
–20
–30
–40
model cluster, dqSi = 0.021 electrons for the root mean square (rms) charge fluc-
tuation in a-Si and dqC = 0.015 electrons in da-C have been obtained. A revised
analysis of a core-level spectroscopic measurement (Ley et al., 1982) yielded the
estimation that the charge fluctuation in a-Si must be lower than 0.030 electrons.
Recently, a similar theoretical model was reported for distorted long disordered
3.1 Classification of semiconductors 27
Figure 3.4. Comparison of Se atomic net charges calculated using the Hartree–
Fock ab initio calculation and eqn. (3.4). Each symbol represents an atom.
(From Lukács and Kugler (2010). Chem. Phys. Lett., 494, 287. Copyright
2013 with permission from Elsevier.)
selenium chains (Lukács and Kugler, 2010). The net charge on the atoms of a
triad depends on the deviation of θ from the average value (dθ = θ – 101°):
qM = 2 Adθ + Bdθ 2 and qK = qL = − Adθ + Bdθ 2 . (3.3)
total
The total atomic net charge on atom M, qM , is a sum of the contributions
originating from the three triads containing atom M:
2
total
qM = 2 AdθM + BdθM 2
− A (dθM−1 + dθM+1 ) − B dθM−1 + dθM+1
2
.
(3.4)
Heat capacity
In the crystalline case, the heat capacity is easy to derive from the lattice vibra-
tion model. A high-temperature limit applying both classical and quantum-
mechanical models provides us with a value of C = 3NkB , where N and kB are
the number of atoms and the Boltzmann constant, respectively. Low-temperature
heat capacity using a quantum-mechanical calculation is in good agreement with
experiments, whereas the classical counterpart (C = 3NkB , the Dulong–Petit
28 Structure
law) is not applicable. The total energy due to the lattice vibration is obtained
by integrating the energy of a single quantum oscillator as follows:
∞
h̄ω
Eosc = h̄ω/2 + g (ω) dω, (3.5)
eh̄ω/kB T −1
0
where ω is the angular frequency (h̄ω is the phonon energy) and g(ω) represents
the vibration density of states per unit frequency range, i.e. the number of modes
with frequency [ω; ω + dω] is g(ω)dω. Only lattice modes of low frequency
will be excited at low temperatures. The dispersion relation has three acoustic
branches, one longitudinal and two transverse:
V ω2 1 2
g(ω) = 3
+ 3 , (3.6)
2π 2 vL vT
where vL and vT are the sound velocities of two different modes. The heat
capacity due to phonon excitation at low temperature is given by
dEosc 2V π 2 kB 1 2 kB T
C= = + ∼ T 3. (3.7)
dT 15 vL vT h̄
In addition, the contribution due to the electrons must be considered for a
more accurate calculation of heat capacity. The most general model is the free
Fermi-electron-gas description. The low-temperature approach obtained from a
Bethe–Sommerfeld expansion yields
π2 T
Cel = NkB ∼ T, (3.8)
2 TF
where TF is the Fermi temperature. Therefore the total heat capacity of crystalline
materials can be written as
Ctotal = γ T + βT 3 . (3.9)
This model does not work for amorphous semiconductors; there is no low-
frequency collective excitation in amorphous materials. Another problem is that
the free electron model is applicable for metals, but it does not work well for
semiconductors. Near linear heat capacity has been observed at low temperatures
for a large class of different materials, including chalcogenide glasses. The
common property among these structures is the low average coordination number
of atoms. In the same year, Phillips (1972) and Anderson, Halperin, and Varma
(1972) proposed the “two-level system” model for the description of linear
dependence of low-temperature heat capacity. The starting point is a fused
3.1 Classification of semiconductors 29
where E is the energy difference between the two states. The heat capacity
can be calculated using
∂ 2F
CV = −T . (3.11)
∂T 2 V
Glass-transition temperature
As discussed in Chapter 2, glasses prepared by the melt-quenched (MQ) method
undergo a glass transition at Tg , which is defined as the transforming temperature
between the glassy and supercooled liquid states. A supercooled liquid describes
a material that remains in a liquid state, even below Tm . Although the definition
of the glass transition itself has a clear meaning, the experimental determination
of Tg is not easy, because the deduced Tg depends on experimental conditions,
such as the cooling rate of the melt in preparation and/or the heating rate in
the DSC measurement itself. It is still unknown which factors dominate Tg
(Elliott, 1990; Greaves and Sen, 2007; Tanaka and Shimakawa, 2011). In spite
of this ambiguity in the determination of Tg , it does correlate to some physical
parameters. A well-known empirical rule is Tg 2Tm 3 (where the temperatures
are measured in kelvin) (Kauzmann, 1948). Recently, it has been suggested that
the quantum effect plays a role in the glass-transition temperature (Novikov and
Sokolov, 2013): quantum effects lead to a significant decrease in Tg with respect
to Tm , so that the ratio Tg Tm can be much smaller than 23 in materials where
Tg is near or below 60 K (lower temperatures), and is given by
Tg A
= , (3.15)
Tm 1 + B/Tg
where A 23 and B depends on the strength of the Boson peak and some other
parameters.
As Tm can be primarily related to a cohesive energy, Tg should also be
related to the cohesive (or bond) energy in covalent glasses; this is discussed
in the following. At the glass-transition temperature, a glassy matrix may be
destroyed; Tg is therefore expected to be related to a network constraint (or
coordination number). Here, we introduce some representative models on this
topic, although there are numerous approaches (see, for example, Tichý and
Tichá, 1995). Tanaka (1985) has proposed the following empirical relation for
3.1 Classification of semiconductors 31
Tg (K)
E –0.9 (eV)
Viscosity
Glasses experience a supercooled liquid state when they are cooled below
their melting points. Glass-forming liquids have high viscosities even at high
3.2 Correlation functions 33
temperatures, and the viscosity of these liquids changes rapidly near the
glass-transition temperature. Thus the temperature at which the viscosity reaches
a certain value (1012 Pa s) is also used as a determination of Tg . The empirical
rule for temperature-dependent viscosity η, called the Vogel–Tammann–Fulcher
equation, in various glasses is given as η = η0 exp[A(T – T0 )], where A and
T0 are constants (Angell, 1988, 1995); T0 is related to the fragility of glasses.
Strong glasses, for example SiO2 , have T0 = 0 (following an Arrhenius relation),
and the fragile structure shows non-Arrhenius behavior. Note that the viscos-
ity is continuous through Tg , whereas the heat capacity shows a discontinuity
(remember the DSC measurement). The above form of the relation appears in
many transport phenomena, such as ionic conduction in liquid states. We will
see this empirical law in Section 3.4 on phase-change phenomena.
The model proposed by Adam and Gibbs (1965) is widely accepted in explain-
ing temperature-dependent viscosity, in which the important factor determining
the relaxation phenomena is the configurational entropy (CE). Alternatively,
Aniya (2002) has suggested that fluctuations in the coordination number Z
and the binding energy E in a glassy system dominate the fragility of glass:
large fluctuations in Z and E produce an increased fragility.
Peak position
(nm) Z
1 0.235 4
2 0.384 12
3 0.45 12
4 0.543 6
5 0.591 12
6 0.665 24
7 0.705 16
8 0.768 12
9 0.803 24
10 0.858 24
11 0.89 12
12 0.94 8
13 0.969 24
Figure 3.7. Schematic for the pair correlation function. Vector r indicates the
direction along which the atomic distribution is calculated.
to the probability of finding a particle inside a volume dV (see Figure 3.7). The
radial distribution function (RDF) is also a measure of the probability of finding
an atom a distance r away from a reference particle. This function, J(r), represents
the number of atoms within a distance between r and r + dr from a given
particle (see Figure 3.8). In contrast to structures that have translation symmetry,
amorphous materials are isotropic; therefore the pair correlation function and
the RDF contain the same information because of the following relationship:
J(r)dV = g(r)4π r2 dr. Functions must be zero below a given distance, which
is related to the atomic repulsion, and they should exhibit alternating maxima
and minima following the coordination shells. The correlation with the atom
3.2 Correlation functions 35
at the origin is lost at large r. It follows that g(r) tends to unity (or to the
density, depending on the normalization), whereas the RDF has a parabolic
shape when r goes to infinity. Scientists working with amorphous materials
usually use J(r). The peak positions in J(r) provide the radii of the different
coordination shells of atoms surrounding the average atom. The area under
the peaks derives the coordination numbers of the different shells of atoms.
In a structural model, the RDF may be easily determined by calculating the
distance between all the atomic pairs and putting them into a histogram. An
alternative form of the RDF that is sometimes used is T(r), which is equal
to J(r)r.
The structures of amorphous semiconductors are typically determined by
neutron, electron, or x-ray diffraction experiments. Neutrons, having no charge,
interact with the nucleus of each atom, whereas x-rays or electrons are sensitive
to the electron distribution inside the amorphous materials. A neutron beam is
capable of penetrating well beyond the surface of a sample, to depths on the
order of centimeters in the condensed phases. In a diffraction experiment, we
measure the intensity as a function of the scattering angle. Thermal neutrons
with wavelengths of 0.1– 0.2 nm are very useful for atomic-scale structural
(and magnetic structure) investigations. Neutron diffraction is a non-destructive
technique; it is a form of elastic scattering in which the neutrons exiting the
sample have the same energy as the incident neutrons. In a standard neutron-
diffraction measurement, g(r) and J(r) are derived from the measured scattered
intensity by Fourier transformation of the static structure factor S(Q), where
Q = kinitial – kfinal . Here, kinitial and kfinal represent the wave vectors in the initial
and final states, and
1
ρ (g (r) − 1) = d Q (S ( Q) − 1) e−i Qr . (3.16)
(2π )3
36 Structure
4
3.5
3
2.5
g (r)
2
1.5
1
0.5
0
0 2 4 6 8 10
r (Å)
Figure 3.9. Pair correlation function of graphite-like a-C measured by neutron
diffraction and derived from S(Q) by Fourier transformation.
where Q = (4π λ)sin θ and r are already scalar values, λ is the de Broglie
wavelength of the neutron, and 2θ is the scattering angle. A special example for
a-C films is shown in Figure 3.9.
Neutron cross-sections can be divided into two parts: coherent and incoherent
cross-sections, σ coh and σ inc , respectively. The elastic part of the coherent scat-
tering provides information about the structure. Measured cross-sections for any
element of the periodic table can be found in Sears (1992), pp. 29–37, or at
http://www.ncnr.nist.gov/resources/n-lengths. The coherent, incoherent, and
absorption cross-sections of isotopes are given in units of barn (1 barn =
10–28 m2 ). In monatomic structure investigation, the absorption cross-section
has no role. Materials with a larger coherent cross-section than their incoherent
counterparts provide a good opportunity to derive the RDF experimentally using
neutron diffraction.
X-rays are scattered by the electrons in atoms. When a covalent bond is
formed, the associated valence electrons are distributed between the two atoms.
It follows that the x-ray diffraction pattern probes the electron distribution around
each atom and between the atoms. In amorphous cases, electrons are placed
asymmetrically, so x-ray scattering may provide a more complex and different
density distribution than that produced by neutron diffraction.
3.2 Correlation functions 37
differences due to the variation of the temperature, but there was no significant
difference between the radial distribution functions.
The structure of another a-C sample (20–30 mg, a very limited specimen mass
for neutron diffraction) prepared by plasma-arc deposition has been determined
by neutron diffraction at the twin-axis diffractometer, D4, at the Institut Laue-
Langevin (ILL), Grenoble, France (Gaskell et al., 1991). The conclusion is in
opposition to the above-mentioned results: the structure factor and reduced radial
distribution functions are similar to those for a-Si and a-Ge, indicating a high
proportion of tetrahedral bonding. The first three main interatomic distances
are very close to bond lengths in diamond crystal (0.152 nm and 0.153 nm;
0.250 nm and 0.252 nm; 0.296 nm and 0.296 nm). A fit to the data yields about
86% tetrahedral bonding.
Thick a-C films were prepared in a high-vacuum system of base pressure
around 10–7 Torr, by rf sputtering on liquid-nitrogen-cooled Cu substrates.
Raman measurements were performed to confirm the amorphous nature of the
deposited films. A 0.8 g sample was used for neutron-diffraction measure-
ments performed at 300 K in the Special Environment Powder Diffractometer
of the Argonne National Laboratory’s Intense Pulsed Neutron Source, IL, USA
(Li and Lannin, 1990). The films were removed from the substrate by using
dilute hydrochloric acid. The RDF was compared to theoretical models, and was
found to exhibit qualitative agreement with a number of basic models, indicating
predominant threefold bonding. The absence of a specific peak at the graphite
intrahexagon distance (third-neighbor peak) indicates that structural models with
such intermediate-range correlations are not correct for a-C. This distinguishes
a-C from local two dimensionally ordered graphite-like materials.
Atomic interactions
Classical empirical potentials
In order to prepare three-dimensional atomic structures of amorphous semi-
conductors, a transferable atomic interaction is required. The simplest ver-
sions of atomic interactions are the classical empirical potentials, which are
42 Structure
Keating potential
The Keating empirical potential (Keating, 1966) is a simple example for a three-
atom local interaction which has macroscopically measurable parameters. It was
originally designed to describe the defect-free diamond type of crystal, such as
silicon or germanium, or diamond-structure carbon, in which the atoms have
four nearest neighbors. The bond-stretching and bond-bending force constants,
α and β, respectively, are determined from the fitting to the elastic properties of
the crystal phase. The number of bonds is equal to 4N2 = 2N, and the number
of atomic triads is equal to (4N × 3)2 = 6N, where N atoms form the geometric
structure. For these materials, the bond length in an equilibrium state is equal
to r0 (0.155 nm, 0.235 nm, and 0.245 nm are for C, Si, and Ge, respectively),
and the bond angle is the tetrahedral angle (Θ = arccos(−13) = 109.47°). In
the Keating potential, bonding geometries which deviate from these equilibrium
values are assigned an energy penalty:
⎡ ⎤
3 ⎣α 2
2N 6N
rij rik + r02 /3 ⎦ ,
2 2
U= 2 r − r02 + β (3.18)
8r0 2 ij ij j ik
where rij is the vector pointing from atom i to atom j, and rij rik is the scalar
product of two vectors. This formula is practically a Taylor expansion where
the linear terms vanish and only the quadratic terms are kept. The constants
α and β are obtained from a fit to the elastic properties of the crystal. For
silicon, values of α = 296.5 eV nm−2 and β = 0.285α were obtained. The
validity of a large deviation from the equilibrium values of bond length and
bond angle is questionable. It is suitable only for small displacements from the
3.3 Three-dimensional structure 43
ideal positions. In particular, it should not be used for very disordered systems.
Another disadvantage is that the adjacency matrix of a topological structure must
be known for the total energy calculation. If the adjacency matrix is unknown,
we cannot calculate the diffusion or other dynamical processes.
Several modified versions of the Keating potential, eqn. (3.18), have been
developed. Fullerenes, nanotubes, and graphenes are formed by threefold-
coordinated carbon atoms. Overney, Zhong, and Tománek (1993) reparam-
eterized the potential, i.e. the bond-stretching and bond-bending force constants
(α, β) were determined by applying the density-functional-theory (DFT) ab ini-
tio local-density approximation total energy calculations for specific distortion
modes of graphite. The bond angle is equal to Θ = 120°, and instead of the
last term, r02 /3, we must substitute the term r02 /2. The structural rigidity and
the low-frequency vibration modes of carbon have been calculated using this
modified version of the Keating potential.
Rücker and Methfessel (1995) worked out a generalization of the Keating
model for group IV semiconductors and their alloys. First, the energy of mixing
relative to the pure materials is separated into two terms, namely Uchemical and
Ustrain . The chemical term arises because the strength of an A–B bond is different
from the average of the A–A and B–B bond strengths. Secondly, higher-order
anharmonic terms are included in order to describe strongly distorted systems.
Even for group III–V alloys (Alx Ga1−x As alloys), a modified Keating potential
has been presented in the literature (Sim et al., 2005). The anharmonic model
potential consists of Coulomb interaction terms in the formulas because of the
charge transfer between different types of atoms.
Stillinger–Weber potential
Generally, any empirical potential function could be described by one-, two-,
three-body, etc., contributions. If there is no external force, this expansion begins
with the pair-wise terms. For sigma-bonded materials the three-body interaction
term at least is required. The Stillinger–Weber empirical potential (Stillinger
and Weber, 1985) that has been developed for liquid and crystalline phases of
silicon is based on two- and three-body interactions:
⎧ ⎫
⎨ ⎬
(2) λ (3)
U = Aε vij rij + v rij , rik , (3.19)
⎩ A j ik j ik ⎭
ij
44 Structure
where ε = 2.16826 eV. The summations include all pairs ij and all triples jik of
atoms in the system. For rij /σ < a, the two-body term is chosen to be
r
rij −p rij −q ij
−1
vij(2) rij = B − exp −a , (3.20)
σ σ σ
otherwise vij(2) (rij ) = 0. This definition automatically cuts off at rij /σ = a with-
out discontinuities in any rij /σ derivatives (σ = 0.20951 nm). If rij /σ and rik /σ
are smaller than the cut-off distance a, the three-body part is given by
−1 r −1
(3) rij ik 2
vj ik rij , rik = exp γ −1 + −1 cos θj ik + 1/3 ,
σ σ
(3.21)
otherwise vj(3)ik (rij , rik ) = 0. In practice, only a small subset of these fragments
should actually be computed because atoms interact only when their separation
is less thanσ a. The most “satisfactory” parameter set is the following:
A = 7.049556277, B = 0.6022245584, p = 4, q = 0,
a = 1.80, λ = 21.0, and γ = 1.20.
Tersoff potentials
A large family of empirical potentials for silicon, carbon, and multicomponent
covalent systems was developed by J. Tersoff (1986, 1988a, 1988b, 1988c).
Sometimes they are named Abell–Tersoff potentials in order to recall the pion-
eering work on empirical pseudopotential theory of G.C. Abell (1985). The pair
potential
1
U= f (rij ) Aexp −λ1 rij − Bij exp(−λ2 rij ) , (3.23)
2 i,j =i
where rij is the distance between atoms i and j, and A, B, λ1 , and λ2 are all
positive; f (rij ) is a cut-off function to restrict the range of the potential. The first
term is repulsive, while the second is interpreted as representing the bonding in
a Morse-type potential. The parameter Bij implicitly includes the bond order
and depends upon local environments. Deviations from the simple pair potential
are ascribed to the dependence of Bij upon the local environment. The bonding
strength Bij should be a monotonically decreasing function of the number of
competing bonds, the strength of the bonds, and the cosines of the angles with
the competing bonds.
46 Structure
Oligschleger potential
The other important class of amorphous semiconductors is the chalcogenides.
The model element for these materials is selenium. An empirical three-body
potential has been developed by Oligschleger et al. (1996) with the intent of
providing both a realistic and a simple description of a selenium–selenium
interaction. The most common selenium crystals are built up by either selenium
rings or infinite helical chains; in both structures every selenium atom has
two nearest neighbors, and this unique property has to be reproduced by the
interatomic potential. The authors have taken into account both small selenium
clusters and crystalline phases during the parameter fitting of the potential. Its
analytical form is given by
U= V2 rij + h rij , rkj , θij k + cyclic permutations. (3.24)
i<j i<j <k
V2,3 (r) = a2,3 exp (αr) + b2,3 exp (βr) + c2,3 exp (γ r) . (3.26)
Otherwise, V2,3 (r) = 0. Note that in Table I in the original paper (Oligschleger
et al. (1996), p. 6165) the parameter a = 9281.2 in the two-body case. The other
misprint is that α = −7.9284 instead of −7.984. Correct parameters in reduced
units (see the original paper) can be taken from Table 3.2.
3.3 Three-dimensional structure 47
Tight-binding models
Tight-binding (TB) models offer a semiempirical quantum chemical approach to
describing the interaction between atoms. The TB model is useful in generating
and describing the atomic structure of amorphous materials. We focus only
on orthogonal TB models that have been developed for carbon, silicon, and
selenium. The total energy of the system is written as follows:
and Erep is the short-range repulsive energy. The electronic eigenvalues are
obtained by diagonalizing an empirical tight-binding Hamiltonian. The on-site
elements are the atomic orbital energies of the corresponding atoms, whereas
48 Structure
and
mc mc
r d0
(r) = 0 (d0 /r) exp m −
m
+ , (3.32)
dc dc
(a)
(b)
Figure 3.10. Bond transposition process for a-Si (a-Ge) structural model con-
structed by Wooten et al. (1985) using MC simulation. (a) Configuration of
bonds in diamond structure; (b) modified configuration after bond transposition.
argon were presented by the authors. This inverse problem solution method is
convenient for the analysis of diffraction data and at the same time for modeling
the amorphous structures. The RMC simulation is based on the results of diffrac-
tion measurements and is free from the description of the interaction between
atoms. This is the only method generating three-dimensional particle configura-
tions that are consistent with the experimentally derived structure factor or the
radial distribution function.
Theoretically, both functions, the structure factor and the RDF, can comprise
the initial data for RMC simulation, but use of S(Q) has an important advantage.
The main problem that occurs at any diffraction measurement – the measure-
ment of Q in only a limited interval and Fourier transformation in an infinite
interval – is bypassed. To speed up the simulation and/or to reach more realistic
configurations, constraint(s) could usually be applied. To start the RMC simu-
lations, a given number of particles are confined in a well-defined box. The
simulation derives the density that is the most important constraint. Another
constraint that is used several times is the coordination number. It can easily be
prescribed that particles must strictly have a given number of particles within a
fixed distance. A hard core diameter (the lower limit of bond length) can also
be applied in order to avoid physically meaningless configurations. For sys-
tems where atomic interactions are covalent, i.e. bonds are directed, introducing
constraints on bond angles might also be a useful constraint.
The use of constraints modifies the original RMC algorithm in the following
way.
(a) Start with an initial particle configuration. Calculate its RDF or structure
factor. Calculate also the difference between the simulated and experimental
RDFs or structure factors (χ02 ). The comparison with experiment is quantified
(yexp −ysim )2
using a function of the form χ 2 = σ2
, where yexp and ysim are the
measured and calculated quantities; σ is a measure of the accuracy of the
measurement. The sum includes all points in a function such as the RDF or
structure factor.
(b) A new, trial, configuration is generated by the random motion of a particle.
(c) Check whether the new configuration satisfies the constraint(s) applied. If
not, start again from (b). (This is an additional step to the standard RMC
simulation.)
(d) Provided that the constraints are satisfied, calculate χn2 for the new trial
configuration.
3.3 Three-dimensional structure 53
2
(e) If χn2 < χ(n−1) , the new configuration becomes the starting configuration, i.e.
the move is accepted; otherwise it is accepted with a probability that follows
the normal distribution.
(f) Repeat the process from (b) until χn2 converges to its “equilibrium” value.
In this way, sets of particle coordinates are generated which are con-
sistent with a given diffraction data set and with the constraint(s) applied.
Figure 3.11 displays three stages of a RMC calculation in two dimensions.
Figure 3.11(a) shows the initial crystalline configuration (right), and the meas-
ured (thick solid line) and calculated (thin solid line) pair correlation func-
tions (left). Figure 3.11(b) displays a middle stage of the RMC simulation, and
Figure 3.11(c) represents a stage where the fit is near optimum (L. Pusztai,
private communication). Free RMC computer code is available at http://www.
szfki.hu/~nphys/rmc++/opening.html. A brief description of the computer code
can be obtained from Gereben et al. (2007) and Gereben and Pusztai (2012).
Constraints
In order to derive constraints, the following question should be answered: what
natural local configurations can be found in nature? During prepartion of the
amorphous phase, the covalently bonded atoms tend to seek out one of the dif-
ferent local configurations. Information about these local orders of the amorph–
ous structures might be provided by analyzing the embedded fragments inside
different large molecules. The environment might be considered as a “white
noise” around the fragments. The Cambridge Structural Database (CSD) (Allen,
Kennard, and Taylor, 1983), the world’s largest experimentally determined crys-
tal structure, has within it the results of x-ray and neutron-diffraction studies.
The CSD is designed as a critically evaluated numerical resource, containing
3D atomic coordinates. It provides a good opportunity to obtain measured lower
and upper limits for bond lengths, bond angles, second-neighbor distances, etc.,
if the CSD contains a large enough number of diffraction records.
A systematic search of structural data can be carried out for X–X and X–X–X
fragments on this database, where X denotes the element forming the disordered
system. We have collected the experimentally determined structural data of
molecules containing Si–Si–Si fragments (Kugler and Varallyay, 2001) and for
Se–Se–Se (Hegedus and Kugler, 2005). Recently, a new search was made1 ,
(a)
(b)
(c)
and Figures 3.12 and 3.13 display the results of these bond-angle distributions
for a-Si and a-Se, respectively. The number of Si–Si–Si fragments is equal to
8071, whereas for Se–Se–Se it is less (905 fragments). The results of a search
for a-Si are surprising. The overwhelming majority of these points fall in the
expected region, i.e. around 0.235 nm and 109.47°. The minimum bond length
3.3 Three-dimensional structure 55
0.35
0.3
0.25
Frequency
0.2
0.15
0.1
0.05
0
40 50 60 70 80 90 100 110 120 130 140 150 160 170 180
Angle (degree)
0.35
0.3
0.25
Frequency
0.2
0.15
0.1
0.05
0
40 50 60 70 80 90 100 110 120 130 140 150 160 170 180
Angle (degree)
Figure 3.13. Histogram of bond-angle distribution of Se–Se–Se fragments col-
lated from the CSD.
is about 0.22 nm, whereas the maximum is 0.27 nm. The average bond length
and angle are 0.237 nm and 106.3°, respectively. There are some extrema.
Two well-defined, unexpected regions can also be found, for example there are
some significant angles in an interval of 75°–96°. These angles belong to the
56 Structure
near planar square arrangement. Several squares were found in the database.
Bond angles around 60° can be also found. The second conclusion of bond-
angle analysis is that nearly equilateral triangles of Si3 are present among the
fragments. This is the other unexpected region, and most of the a-Si theoretical
models do not contain this type of structure, i.e. three- and four-membered rings.
A similar conclusion can be delivered for a-Se.
1.0
(a)
0.5
S(Q)–1
0.0
–0.5
–1.0
0 5 10 15 20 25
Q (Å–1)
1.0
(b)
0.5
S(Q)–1
0.0
–0.5
–1.0
0 5 10 15 20 25
Q (Å–1)
Figure 3.14. (a) Comparison of the RMC solution for Model1 and the experi-
mental S(Q) for a-Si. (b) The same comparison for Model3 for a-Si. (Taken
with permission from Kugler et al. (1993). Phys. Rev. B, 48, 7685. Copyright
2013 by the American Physical Society.)
neighbors in the first coordination shell. The upper limit for the bond length
was 0.27 nm. The g(r) function obtained by the Fourier transformation of the
measured S(Q) almost reached its minimum between the first and second main
peaks at 0.27 nm. Furthermore, the CSD suggests a similar value for the max-
imum first-neighbor distance. Model 3 satisfied the most complicated constraint.
In addition to the atoms having four neighbors, if an attempted move resulted
in newly formed bond angles with an average that did not fit into a normal-
like distribution centered on the tetrahedral angle, the move was immediately
rejected. In systems such as a-Si, it is a normal requirement that most of the
angles should be roughly tetrahedral. Angles far from the tetrahedral value are
less probable, but permitted. Figure 3.14(a) shows the comparison of the RMC
solution for Model 1 and the experimental S(Q), and Figure 3.14(b) gives the
same comparison for Model 3. The overall agreement is quite good. It is thought
58 Structure
4 (a)
g(r)
2
0
0 2 4 6 8 10
r (Å)
5
4 (b)
3
g(r)
0
0 2 4 6 8 10
r (Å)
5
4 (c)
3
g(r)
0
0 2 4 6 8 10
r (Å)
Figure 3.15. (a), (b) Pair correlation functions calculated using RMC with dif-
ferent constraints. (c) Pair correlation function obtained by Fourier transforma-
tion of measured S(Q). (Taken with permission from Kugler et al. (1993). Phys.
Rev. B, 48, 7685–7688. Copyright 2013 by the American Physical Society.)
3.3 Three-dimensional structure 59
1.0
0.0
–1.0 –0.5 0.0 0.5 1.0
Cosine of bond angle
Figure 3.16. Cosine of bond-angle distributions for three different models.
(Taken with permission from Kugler et al. (1993). Phys. Rev. B, 48, 7685–
7688. Copyright 2013 by the American Physical Society.)
that the greater amplitude of the oscillations at higher Q-values and the small-
angle-scattering part of the S(Q) curve could be fitted more successfully using a
much bigger box.
The RDF for Model 1 (identical to Model 2) is displayed in Figure 3.15(a),
and that for Model 3 is given in Figure 3.15(b). Figure 3.15(c) depicts g(r)
obtained by using a Fourier transform of the measured S(Q). The characteris-
tics of these functions are rather similar. Cosine distributions of bond angles
(cos θ ) lead to significant differences between microscopic structures obtained
using three different constraints (Figure 3.16). This makes clear that a diffraction
experiment by itself provides insufficient information to determine unambigu-
ously the microscopic local structure. Identical S(Q) and/or g(r) functions can
be consistent with substantially different bond-angle distributions. The cosine
distribution for Model 1 has a large peak at around 60°, and also a large propor-
tion of angles were found near 180°. However, an intense peak around 109.5°
is provided by the model. These results are completely different from those of
the other two models. Model 2 reflects probably the most fundamental require-
ment that atoms in any solid form of Si should be fourfold coordinated. This
simple constraint greatly reduces the freedom of moving atoms around. The
main problem with this picture might be the probable over-representation of
angles higher than 130°. Model 3 was introduced to narrow the cosine distri-
bution further, but no essential narrowing could be achieved compared with
Model 2.
60 Structure
Wooten’s model of a-Si contains internal bond angles between about 90° and
150°. The classical empirical Keating potential was applied for the construction
in which the interaction has a quadratic energy term for the difference between
the cosine bond angle and the cosine ideal bond angle. This term avoids the large
deviation from the canonical value of bond angle. Note that the RMC simulation
does not take the energy into account. The existence of angles smaller than
90° or 60° also agrees with the CSD. A Japanese group investigating a-Si:H by
RMC simulation also concluded the existence of small bond angles (Tabuchi
et al., 2004). Furthermore, another Japanese group (Ishimaru, Yamaguchi, and
Hirotsu, 2004) investigated silicon–germanium alloys by a MD method using a
classical empirical potential of Tersoff (Tersoff, 1989), and they also observed
bond angles of 60°.
A comparison between the experimental structure factors of evaporated and
ion-implanted a-Si has been modeled by the RMC simulation (Pusztai and
Kugler, 2005). A detailed comparison, in terms of the pair correlation func-
tion and the cosine distribution of bond angles, has been reported for the two
materials. It is found that for an acceptable RMC reproduction of the measured
structure factors, the evaporated models must contain more “small” bond angles
than their necessary abundance in the implanted models.
g(r) 2
0
1 2 3 4 5 6
r (Å)
Figure 3.17. Pair correlation functions obtained by RMC simulation applying
different constraints. Solid line: hard-sphere constraints only; dashed line: long
chains; and dots: Se8 rings. (Taken with permission from Jóvári, Delaplane,
and Pusztai (2003). Phys. Rev. B, 67, 172201. Copyright 2013 by the American
Physical Society.)
ball-milled a-Se sample. The obtained structure factor was interpreted by RMC
simulation. Selenium has several crystalline forms, consisting of Se8 rings or
chains. The authors concluded from the diffraction data that it is not possible
to determine whether a-Se consists of long chains, Se8 rings, or a mixture of
both phases. It is possible to construct both of the models whenever the structure
factors agreed quantitatively with measured data (Figure 3.17). The measured
sample was made from crystalline Se consisting of long chains; it is suggested
that the investigated sample contained primarily chain-like structures.
Molecular-dynamics simulation
An alternative 3D structure model construction is the deterministic molecular-
dynamics (MD) method for describing non-crystalline systems. The primary
application of MD is to follow the preparation process in situ and to reach a
condensed phase. In this process, all that is assumed is the validity of classical
mechanics. This approximation is never truly invalid. An indicator of the valid-
√
ity of classical mechanics is the de Broglie wavelength, λ = h/ 2mE. For a
particle moving with a kinetic energy of 1 eV, the de Broglie wavelength of
62 Structure
and at the earlier time (t – dt), and (ii) V(r), we can derive the new configuration
at time (t + dt). Knowledge of r (t − dt) provides a little difficulty at the first
MD step, but this is easy to handle as long as we know the initial velocity.
Furthermore, we can calculate the atomic velocities at time t, which is important
for the temperature estimation. The local error in a given position is proportional
to O(dt 4 ). If we calculate the cumulative error for the positions over a given
time period T, we obtain
There are other methods for the numerical integration of ordinary differential
equations of order 2, such as Beeman’s algorithm (Beeman, 1976), or the well-
known Runge–Kutta method, but in most cases the velocity Verlet algorithm is
applied.
Temperature control
In an amorphous sample, the preparation temperature plays a crucial role. The
temperature must be controlled during quenching from the liquid state and in
the deposition to a substrate. Most of the main preparation techniques utilize
low-temperature-induced amorphization. Each of these preparation techniques
is highly non-equilibrium in nature. A heat bath removes energy from the sample
during the amorphization. At every MD simulation step the temperature must be
monitored. The instantaneous temperature is given by the following well-known
statistical mechanics equation:
3 1N
NkB T = pi2 . (3.43)
2 i=1
2m
To simulate the heat conduction outside the simulation box, and to bring the
structure back to the desired temperature, one is required somehow to control the
temperature. During the process of film growth, particles bombard the growing
film. Deposition to a substrate requires that the substrate temperature is kept
constant. The instantaneous kinetic energies of the substrate atoms (and the
temperature of the substrate), however, may increase during MD simulations due
to particle impact, and they are dependent on the energy of the incoming particle.
At rapid cooling, the characteristic rate of quenching is around 106 K s−1 , or
lower in experiments. A variety of thermostat methods used to control the
temperature can be found in literature, but the Nosé–Hoover thermostat (Nose,
1984; Hoover, 1985; Chen, Wu, and Cheng, 2011) and the simpler velocity-
rescaling methods are the most popular. Equations describing the Nosé–Hoover
thermostat have a free parameter, and sometimes its derivation is difficult.
A MD computer code (ATOMDEP) has been developed to simulate the
preparation procedure of amorphous semiconductors, which are usually grown
by a vapor-deposition technique on a substrate (Kohary and Kugler, 2001). It
should be noted that experimentally no rapid quenching preparation method has
been reported for group IV amorphous materials. To integrate the MD equations
of motion, the velocity Verlet algorithm was included with a time step equal to
3.3 Three-dimensional structure 65
about 1 fs. The substrate temperature was fixed by velocity rescaling. During the
growth procedure, an atom started to move toward the target surface. In order
to avoid surface effects, periodic boundary conditions in two dimensions were
adopted. The initial x, y coordinates and the deposition energies were randomly
distributed. The bottom layer of the substrate was fixed at its ideal lattice site in
order to stabilize mechanically the substrate.
A typical deposition rate during sample preparation in the laboratory is around
10 atoms s−1 cm−2 . The surface of the substrate in computer experiments is
14
around 10–14 cm2 , i.e. one incoming atom per second should reach the sub-
strate. This deposition rate presents a difficulty for the simulation, and it is
impossible to handle easily using MD with today’s computer facilities within
a tight-binding atomic interaction model. Despite this disadvantage, the high
theoretical deposition rate applied is low enough to warrant the relaxation of the
previous atom before the next deposition. This developed computer code was
successfully applied to prepare a-C, a-Si, and a-Se, and this will be discussed in
the following section.
Distance profile
T = 100 K
25 T = 300 K
Density of atoms in z directions
20
15
10
0
–5 –2.5 0 2.5 5 7.5 10 12.5
r (Å)
Figure 3.18. Density profiles of two networks (T = 100 K and T = 300 K)
are displayed perpendicular to the substrate surface. The arrows represent the
layer positions of a perfect diamond crystal. A memory effect is observed near
the surface of the substrate. (Taken with permission from Kohary and Kugler
(2001). Phys. Rev. B, 63, 193404. Copyright 2013 by the American Physical
Society.)
50
J(r)
40
30
20
10
0
1 1.5 2 2.5 3
r (Å)
Figure 3.19. Partial radial distribution functions J(r) of an a-C model con-
structed by TBMD simulation. Numbers following letters C refer to the coor-
dination numbers of carbon atoms. (Taken with permission from Kohary and
Kugler (2001). Phys. Rev. B, 63, 193404. Copyright 2013 by the American
Physical Society.)
thick layers over the [111] surface, the sp3 content is high due to the memory
effect. In the rest of the bulk, the sp2 content is dominant. The partial distribution
functions of a model (T = 100 K, E = 1 eV) are shown in Figure 3.19, where
i and j denote the coordination numbers of the connected atoms. The average
first-neighbor distances are 0.143 nm, 0.152 nm, and 0.159 nm for C3–C3, C3–
C4, and C4–C4, respectively. The average coordination numbers for different
models were between 3.1 and 3.4. Coordination numbers larger than four were
not observed.
Angles are analyzed in detail according to the coordination numbers of their
neighbors. Atoms with sp3 local arrangements have nearly the same average
values in the C3–C4–C3, C3–C4–C4, and C4–C4–C4 cases, which are around
the tetrahedral angle. Clear differences are observed for sp2 configurations. The
C4–C3–C4 angles have much lower average values than the others. The latter
average bond angles are close to 120°. This means that threefold-coordinated
central atoms with at least two C3 first neighbors are always near planar struc-
tures, even if the third one is C4. The ring statistics for our models were also
calculated. In the graphitic and diamond networks, sixfold and eightfold rings
68 Structure
are present. The size of a ring is denoted by the number of atoms in a closed path.
In our models, even-membered as well as odd-membered rings can be found,
i.e. 4–7-membered rings.
There were 108 fixed atoms at the bottom of the substrate. The remaining
atoms could move with full dynamics. The simulation cell was open along the
positive z direction, and periodic boundary conditions were applied in the x and y
directions. The kinetic energy of the atoms in the substrate was rescaled at every
MD step (dt = 1 fs) to keep the substrate at a constant temperature of 100 K.
In the deposition process the frequency of the atomic injection was 300 fs−1
(Figure 3.21).
Three different structures were constructed, with average bombarding ener-
gies of 0.1 eV, 1 eV, and 10 eV. At the end of the deposition, atomic networks
contained about 1000 selenium atoms. Rapid cooling is frequently applied
to construct glassy structures. The system is usually cooled down to room
70 Structure
Figure 3.21. Part of the a-Se structure grown with 1 eV bombarding energy at
100 K. There are 324 substrate atoms at the bottom, of which 108 were held
fixed to mimic the bulk crystal. (From Hegedus, Kohary, and Kugler (2004a).
J. Non-Cryst. Solids, 338–340, 283. Copyright 2013 with permission from
Elsevier.)
y
Figure 3.22. One of the several selenium glassy networks constructed using
the rapid cooling technique. It contains 162 Se atoms with periodic boundary
condition in the x and y directions. (From Hegedus et al. (2004b). J. Non-Cryst.
Solids, 338–340, 557. Copyright 2013 with permission from Elsevier.)
than the deposited counterpart, and is an observable difference between the two
preparation techniques.
Secondly, for the description of the interaction between the selenium atoms,
we used the TB model developed by Molina and co-workers (Molina et al.,
1999). Self-consistency was taken into account via the usual on-site Hubbard
term, and was found to reduce any large charge transfer (Lomba, Molina, and
Alvarez, 2000). The time step was equal to 2 fs. The temperature was controlled
via the velocity-rescaling method. Several different glassy selenium networks
were prepared in a rectangular box with periodic boundary conditions; samples
contained 162 atoms (Figure 3.22). The “cook and quench” sample preparation
is described as follows.
First, the temperature of the system was kept at 5000 K for the first 300 MD
steps to randomize the atomic positions. During the next 2200 MD steps, the
temperature was decreased linearly from 700 K to 250 K, driving the sample
through the glass transition and reaching the condensed phase. Then the final
temperature of 20 K was reached and the sample for 500 MD steps (1 ps)
was relaxed. The periodic boundary conditions were lifted along the z direc-
tion at this point. This procedure provided a slab geometry containing periodic
boundary conditions in only two dimensions. The system was then relaxed
for another 40 000 MD steps (80 ps) at 20 K. Short quenching times in the
simulation compared to those in the experiments might lead to many liquid-
state defects being retained in the amorphous structure. Therefore, the Hubbard
72 Structure
25
15
10
0
2.0 2.5 3.0 3.5
°
Distance (A)
parameter U was taken to be 5 eV for the first 4000 MD steps during quenching to
avoid a large number of coordination defects, especially onefold- and threefold-
coordinated atoms. Then U was changed to its accepted value of 0.875 eV for
selenium. The RDF for one of the selenium glassy networks at 20 K is shown in
Figure 3.23. The first main peak at 0.24 nm belongs to covalently bonded
atoms. The crystalline nearest-neighbor distance is 0.237 nm. The second peak
at 0.36 nm corresponds to the intrachain second-nearest-neighbor distance as
expected. The prepeak at 0.33 nm reflects the smallest interchain atomic dis-
tances in a-Se. In simulations at higher temperatures (T = 300 K), these two
peaks merge, as observed in Figure 3.23.
Although simulations based on empirical potentials or TB models have
achieved remarkable success, extension of the simulation methods to more ac-
curate quantum mechanics would be desirable. Unfortunately, DFT within
the most popular local-density approximation (LDA) is very computer time
consuming – it is difficult to perform a separate self-consistent electronic mini-
malization at every MD step. An approach known as the Car–Parrinello method
(Car and Parrinello, 1985) overcomes this difficulty, and it has been successfully
applied to MD simulations of covalently bonded amorphous semiconductors,
yielding valuable structural information for small systems. Structural simula-
tions were carried out on the basis of the Car–Parinello method for a-C (Galli,
Martin, Car, and Parrinello, 1989), a-C:H (Iarloni, Galli, and Martini, 1994),
ta-C (Marks et al., 1996), a-Si (Stich et al., 1991), and a-Si:H having far fewer
than 100 atoms (Buda, Chiarotti, Car, and Parrinello, 1991).
3.4 Phase change and its applications 73
Crystallization kinetics
The dynamics of crystallization in glasses has been fairly well interpreted in
terms of the classical nucleation and growth model developed by Kolmogorov,
Johnson and Mehl, and Avrami (the so-called KJMA model) (see, for ex-
ample, Pineda and Crespo, 1999; Tonchev and Kasap, 2006). The KJMA model
yields the macroscopic evolution of the transformed phase under an isothermal
annealing condition with its simplicity, and it has been extended to analyze
transformations under a constant heating rate. The KJMA formalism may still
present some difficulties, for example fitting the KJMA results to experimental
data produces unreasonable physical parameters for the crystallization, which
have not been well understood so far (Senkader and Wright, 2004).
We briefly discuss now what kind of difficult problems exist in the KJMA
formalism. We consider only the macroscopic properties, without addressing any
specific microscopic view, which is still a matter of debate. These topics will
be described later on. The KJMA theory basically assumes that the nucleation
sites for the new phase are randomly distributed over the volume V, and that the
growth ceases in a region where two neighboring new phase grains impinge on
each other. The transformed volume Vtr is given by
V tr
dV = 1 −
tr
dV ex . (3.44)
V
Here, V ex is called the extended volume, which is the virtual volume of the
particles growing without any impingement. Integration of the above equation
produces the familiar KJMA equation:
f = 1 − exp −f ex , (3.45)
Here the effective growth rate is given by G = drdt, and then f ex at time t is
expressed by
4π 4π
f ex = Nr 3 = N(Gt)3 . (3.47)
3 3
We then obtain the well-known KJMA equation:
1 − f = exp −(kt)n , (3.48)
where C is a constant and D is the fractal dimension of the surface of the grains.
Finally, we obtain f :
1 − f = exp −(kt)D+1 . (3.51)
References
Abell, G.C. (1985). Empirical chemical pseudopotential theory of molecular and metallic
bonding. Phys. Rev. B, 31, 6184–6196.
Adam, G. and Gibbs, J.H. (1965). On the temperature dependence of cooperative relax-
ation properties in glass-forming liquids. J. Phys. Chem., 43, 139–146.
Allen, F.H., Kennard O., and Taylor, R. (1983). Systematic analysis of structural data as
a research technique in organic chemistry. Acc. Chem. Res., 16, 146–153.
Anderson, P.W., Halperin, B.I., and Varma, C.M. (1972). Anomalous low-temperature
thermal properties of glasses and spin glasses. Phil. Mag., 25, 1–9.
Angell, C.A. (1988). Structural instability and relaxation in liquid and glassy phases
near the fragile liquid limit. J. Non-Cryst. Solids, 102, 205–221.
Angell, C.A. (1995). Formation of glasses from liquids and biopolymers. Science, 267,
1924–1935.
Aniya, M. (2002). A model for the fragility of the melts. J. Therm. Anal. Calorim., 69,
971–978.
References 79
Ishimaru, M., Yamaguchi, M., and Hirotsu, Y. (2004). Structural relaxation of amorph-
ous silicon-germanium alloys: molecular-dynamics study. Jpn. J. Appl. Phys., 43,
7966–7970.
Jian, Z., Kaiming, Z., and Xide, X. (1990). Modification of Stillinger-Weber potentials
for Si and Ge. Phys. Rev. B, 41, 12915–12918.
Jóvári, P., Delaplane, R.G., and Pusztai, L. (2003). Structural models of amorphous
selenium. Phys. Rev. B, 67, 172201, 1–4.
Justo, J.F., Bazant, M.Z., Kaxiras, E., Bulatov, V.V., and Yip, S. (1998). Interatomic
potential for silicon defects and disordered phases. Phys. Rev. B, 58, 2539–2550.
Kakinoki, J., Katada, K., Hanawa, T., and Ino, T. (1960). Electron diffraction study of
evaporated carbon film. Acta Cryst., 13, 171–179.
Kauzmann, W. (1948). The nature of glassy state and the behaviour of liquids at low
temperatures. Chem. Rev., 43, 219–256.
Keating, P.N. (1966). Effect of invariance requirements on the elastic strain energy of
crystals with application to the diamond structure. Phys. Rev., 145, 637–645.
Kelires, P.C. and Tersoff, J. (1988). Glassy quasithermal distribution of local geometries
and defects in quenched amorphous silicon. Phys. Rev. Lett., 61, 562–565.
Klug, D.D. and Whalley, E. (1982). Effective charges of amorphous silicon, germanium,
arsenic, and ice. Phys. Rev. B, 25, 5543–5546.
Kohary, K. and Kugler, S. (2001). Growth of amorphous carbon. Low energy molecular
dynamics simulation of atomic bombardment. Phys. Rev. B, 63, 193404, 1–4.
Kohary, K. and Kugler, S. (2004). Growth of amorphous silicon: low energy molecular
dynamics simulation of atomic bombardment. Mol. Simul., 30, 17–22.
Kolobov, A and Tominaga, J. (2012). Chalcogenides: Metastability and Phase Change
Phenomena. Berlin and Heidelberg: Springer.
Kolobov, A.V, Fons, P., Frenkel, A.I., and Ankudinov, A.L. (2004). Understanding of
the phase-change mechanism of rewritable optical media. Nature Mat., 3, 703–708.
Kramer, B., King, H., and Mackinnon, A. (1983). Charge fluctuations in hydrogenated
amorphous silicon. J. Non-Cryst. Solids, 59–60, 73–76.
Kugler, S. and Náray-Szabó, G. (1987). Charge distribution in amorphous silicon clus-
ters: quantum chemical study combined with ring statistics. J. Non-Cryst. Solids,
97–98, 503–506.
Kugler, S. and Náray-Szabó, G. (1991a). Weak bonds and atomic charge distribution in
hydrogenated amorphous silicon. J. Non-Cryst. Solids, 137–138, 295–298.
Kugler, S. and Náray-Szabó, G. (1991b). Atomic charge distribution in diamondlike
amorphous carbon. Jpn. J. Appl. Phys., 30, L1149–L1151.
Kugler, S. and Varallyay, Z. (2001). Possible unusual atomic arrangements in the struc-
ture of amorphous silicon. Phil. Mag. Lett., 81, 569–574.
Kugler, S., Surjan, P.R., and Náray-Szabó, G. (1988). Theoretical estimation of static
charge fluctuation in amorphous silicon. Phys. Rev. B, 37, 9069–9071.
Kugler, S., Molnar, G., Peto, G. et al. (1989). Neutron-diffraction study of the structure
of evaporated pure amorphous silicon. Phys. Rev. B, 40, 8030–8032.
Kugler, S., Pusztai, L., Rosta, L., Chieux, P., and Bellissent, R. (1993a). Structure of
evaporated pure amorphous silicon: neutron-diffraction and reverse Monte Carlo
investigation. Phys. Rev. B, 48, 7685–7688.
Kugler, S., Shimakawa, K., Watanabe, T., Hayashi, K., Laszlo, I., and Bellissent, R.
(1993b). The temperature dependence of amorphous carbon structure. J. Non-Cryst.
Solids, 164–166, 1143–1146.
Kwon, I., Biswas, R., Wang, C.Z., Ho, K.M., and Soukoulis, C.M. (1994). Transferable
tight-binding models for silicon. Phys. Rev. B, 49, 7242–7250.
82 Structure
Laaziri, K., Kycia, S., Roorola, S. et al. (1999). High resolution radial distribution
function of pure amorphous silicon. Phys. Rev. Lett., 82, 3460–3463.
Langer, J.S. (2006). Excitation chains at the glass transition. Phys. Rev. Lett., 97, 11504,
1–4.
Laszlo, I. (1998). Formation of cage-like C60 clusters in molecular dynamics simulations.
Europhys. Lett., 44, 741–746.
Laszlo, I. (1999). Molecular dynamics study of the C60 molecule. J. Mol. Struct.:
THEOCHEM, 463, 181–184.
Lee, T.H. and Elliott, S.R. (2011). Ab initio computer simulation of the early stage of the
crystallization: application to Ge2 Sb2 Te5 phase-change materials. Phys. Rev. Lett.,
107, 145702, 1–5.
Lenovsky, T., Kress, J.D., Kwon, I. et al. (1997). Highly optimized tight-binding model
of silicon. Phys. Rev. B, 55, 1528–1544.
Ley, L., Reichardt, J., and Johnson, R.L. (1982). Static charge fluctuations in amorphous
silicon. Phys. Rev. Lett., 49, 1664–1667.
Li, F. and Lannin, J.S. (1990). Radial distribution function of amorphous carbon. Phys.
Rev. Lett., 65, 1905–1908.
Liu, X.Y. and Sawant, P.D. (2002). Mechanism of the formation of self-organized
microstructures in soft functional materials. Adv. Mater., 14, 421–426.
Lomba, E., Molina, D., and Alvarez, M. (2000). Hubbard corrections in a tight-binding
Hamiltonian for Se: effects on the band structure, local order, and dynamics. Phys.
Rev. B, 61, 9314–9321.
Lukács, R. and Kugler, S. (2010). A simple model for the estimation of charge accumu-
lation in amorphous selenium. Chem. Phys. Lett., 494, 287–288.
McGreevy, R.L. and Pusztai, L. (1988). Reverse Monte Carlo simulation: a new
technique for the determination of disordered structures. Molec. Sim., 1, 359–
367.
Marks, N.A., McKenzie, D.R., Pailthorpe B.A., Bernasconi, M., and Parrinello, M.
(1996). Microscopic structure of tetrahedral amorphous carbon. Phys. Rev. Lett.,
76, 768–771.
Metropolis, N., Rosenbluth, A.W., Rosenbluth, M.N., Teller, A.H., and Teller, E. (1953).
Equation of state calculations by fast computing machines. J. Chem. Phys., 21,
1087.
Molina, D., Lomba, E., and Kahl, G. (1999). Tight-binding model of selenium disordered
phases. Phys. Rev. B, 60, 6372–6382.
Moss, S.C. and Graczyk, J.F. (1969). Evidence of voids within the as-deposited structure
of glassy silicon. Phys. Rev. Lett., 23, 1167–1171.
Nose, S. (1984). A unified formulation of the constant temperature molecular-dynamics
methods. J. Chem. Phys., 81, 511–519.
Novikov, V.N. and Sokolov, A.P. (2013). Role of quantum effects in the glass transition.
Phys. Rev. Lett., 110, 065701, 1–5.
Oligschleger, C., Jones, R.O., Reimann, S.M., and Schober, H.R. (1996). Model
interatomic potential for simulations in selenium. Phys. Rev. B, 35, 6165–
6173.
Overney, G., Zhong, W., and Tománek, D. (1993). Structural rigidity and low frequency
vibrational modes of long carbon tubules. Z. Phys. D, 27, 93–96.
Phillips, J.C. (1973). Bonds and Bands in Semiconductors. New York and London:
Academic Press.
Phillips, W.A. (1972). Tunneling states and the low-temperature thermal expansion of
glasses. J. Low Temp. Phys., 7, 351–360.
References 83
Tang, M.S., Wang, C.Z., Chan, C.T., and Ho, K.M. (1996). Environment-dependent
tight-binding potential model. Phys. Rev. B, 53, 979–982.
Tegze, M. and Faigel, Gy. (1996). X-ray holography with atomic resolution. Nature,
380, 49–51.
Terao, M., Morikawa, T., and Ohta, T. (2009). Electrical phase-change memory: funda-
mentals and state of the art. Jpn. J. Appl. Phys., 48, 080001, 1–14.
Tersoff, J. (1986). New empirical model for the structural properties of silicon. Phys.
Rev. Lett., 56, 632–635.
Tersoff, J. (1988a). New empirical approach for the structure and energy of covalent
systems. Phys. Rev. B, 37, 6991–7000.
Tersoff, J. (1988b). Empirical interatomic potential for silicon with improved elastic
properties. Phys. Rev. B, 38, 9902–9905.
Tersoff, J. (1988c). Empirical interatomic potential for carbon, with application to
amorphous carbon. Phys. Rev. Lett., 61, 2879–2882.
Tersoff, J. (1989). Modeling solid-state chemistry: interatomic potentials for multi-
component systems. Phys. Rev. B, 39, 5566–5568. (Erratum: (1990) Phys. Rev. B,
41, 3248.)
Tichý, L. and Tichá, H. (1995). Covalent bond approach to the glass-transition tempera-
ture of chalcogenide glasses. J. Non-Cryst. Solids, 189, 141–146.
Tonchev, D. and Kasap, S.O. (2006). Thermal properties and thermal analysis: funda-
mentals, experimental techniques and applications. In The Springer Handbook of
Electronic and Photovoltaic Materials, eds. S.O. Kasap and P. Capper. Heidelberg:
Springer, chap. 19.
Tóth, G. and Náray-Szabó, G. (1994). Novel semiempirical method for quantum Monte
Carlo simulation: application to amorphous silicon. J. Chem. Phys., 100, 3742–
3746.
Treacy, M.M.J. and Borisenko, K.B. (2012). The local structure of amorphous silicon.
Science, 335, 950–953.
Valladares, A.A., Alvarez, F., Liu, Z., Sticht, J., and Harris, J. (2001). Ab initio studies
of the atomic and electronic structure of pure and hydrogenated a-Si. Eur. Phys. J.
B, 22, 443–453.
Verlet, L. (1967). Computer “experiments” on classical fluids. I. Thermodynamical
properties of Lennard−Jones molecules. Phys. Rev., 159, 98–103.
Vink, R.L.C., Barkema, G.T., van der Weg, W.F., and Mousseau, N. (2001). Fitting the
Stillinger–Weber potential to amorphous silicon. J. Non-Cryst. Solids, 282, 248–
255.
Walters, J.K., Gilkes, K.W.R., Wicks, J.D., and Newport, R.J. (1998). Progress in mod-
elling the chemical bonding in tetrahedral amorphous carbon. Phys. Rev. B, 58,
8267–8276.
Wang, C.Z., Pan, B.C., and Ho, K.M. (1999). An environment-dependent tight-binding
potential for Si. J. Phys.: Condens. Matter, 11, 2043–2049.
Wilson, M. and Salmon, P.S. (2009). Network topology and the fragility of tetrahedral
glass-forming liquids. Phys. Rev. Lett., 103, 157801, 1–4.
Winer, K. and Cardona, M. (1986). Theory of infrared absorption in amorphous silicon.
Solid State Commun., 60, 207–211.
Wooten, F. and Weaire, D. (1984). Generation of random network models with periodic
boundary conditions. J. Non-Cryst. Solids, 64, 325–334.
Wooten, F., Winer, K., and Weaire, D. (1985). Computer generation of structural models
of amorphous Si and Ge. Phys. Rev. Lett., 54, 1392–1395.
References 85
Xu, C.H., Wang, C.Z., Chan, C.T., and Ho, K.M. (1992). A transferable tight-binding
potential for carbon. J. Phys.: Condens. Matter, 4, 6047–6057.
Yamada, N. (2009). Development of materials for third generation optical strange
media. In Phase Change Materials: Science and Applications, eds. S. Raoux and
M. Wuttig. New York: Springer, chap. 10.
Yang, R. and Singh, J. (1998). Study of the stability of hydrogenated amorphous silicon
using tight-binding molecular dynamics. J. Non-Cryst. Solids, 240, 29–34.
Yelon, A., Movaghar, B., and R.S. Crandall, R.S. (2006). Multi-excitation entropy: its
role in thermodynamics and kinetics. Rep. Prog. Phys., 69, 1145–1194.
4
Electronic structure
Group IV elements
The framework of the linear combination of atomic orbitals (LCAO) can
be used to describe the electronic structure and bonds in these materials.
86
4.1 Chemical bonds 87
Figure 4.1. The sp3 orbitals overlap with the similar orbitals of four adjacent
atoms.
Figure 4.2. An sp3 orbital remains as an unpaired and unsatisfied valence state.
This dangling bond (DB) forms an electronic localized state.
It can be concluded from diffraction measurements that the upper limit for
the first-neighbor distance in a-Si is about 0.27 nm. A special case occurs when
only three neighboring atoms are situated within this distance. Three sp3 orbitals
overlap with the orbitals of neighbor atoms and form normal sigma bonds.
The three bonds are usually referred to as “backbonds” and the three atoms
as “backbond atoms.” One sp3 orbital remains as an unpaired and unsatisfied
valence state, as shown in Figure 4.2. This is called the dangling bond (DB),
and it is localized. This single electron has an uncompensated spin, which can
be detected by electron spin resonance (ESR) measurements. Dangling bonds
dominate electronic and optical properties and hence affect the characteristics of
the device. Amorphous semiconductors with lesser defects (DBs, etc.) are often
referred to as device-quality materials.
be sp2 hybrids, because one pi bond is required for the double bond between the
two carbon atoms. Only three sigma bonds can be formed per carbon atom. In
sp2 hybridization, the 2s orbital is mixed with only two (px and py ) of the three
available 2p orbitals. These three hybrids are also orthonormalized states. The
mathematical terms are as follows:
1/2 1/2
1 2
sp 2 = s+ px ,
3 3
1/2 1/2 1/2
1 1 1
sp 2 = s− px + py , (4.2)
3 6 2
1/2 1/2 1/2
1 1 1
sp 2 = s− px − py .
3 6 2
In ethylene the two carbon atoms form a sigma bond by overlapping two sp2
orbitals, and each carbon atom forms two other covalent bonds with hydrogen
via an s–sp2 overlap, all with 120° angles. The pi bond between the carbon atoms
perpendicular to the molecular plane is formed by the 2pz – 2pz overlap. The
hydrogen–carbon bonds are all of equal strength and length (resonant bonds),
which agrees with experimental data. In a honeycomb graphite crystal, such
a local arrangement can be observed as planar. The bond length is shorter by
0.01 nm, and the bond is stronger than in a diamond crystal because of the
additional pi bond. This local arrangement with sp2 hybridization forms the
graphite-like a-C.
Group VI elements
Selenium atoms contain six valence electrons (4s2 4p4 ), and the previously
described sp3 hybridization is no longer valid. From s, px , py , and pz only
four independent orbitals can be constructed, i.e. two out of four must be
doubly occupied. A simple hydrogen selenide molecule has the atomic struc-
ture H–Se–H, with a bond angle of 91°. Note that in hydrogen telluride the
bond angle is equal to 90°. As a first approximation, we can explain the bond-
ing process in the following way. Two near-perpendicular sigma bonds can
be constructed by putting one electron in the px and another in the py atomic
orbitals, which are overlapped by the 1s orbital of hydrogen atoms. The other
two unshared electrons occupy the pz atomic orbital, forming a non-bonding
90 Electronic structure
Figure 4.3. The px and py orbitals overlap with similar orbitals of adjacent
atoms and form sigma bonds. Two unshared electrons occupy the pz atomic
orbital, forming a non-bonding lone-pair (LP) state.
lone-pair (LP) state. The other two electrons remain untouched in the s atomic
orbital.
Pure non-crystalline selenium is the model material for chalcogenide glasses.
In the condensed phase, selenium atoms form a chain-like structure. It can be
concluded that only two electrons in the px and py orbitals form sigma bonds that
overlap with similar orbitals of adjacent atoms (see Figure 4.3). Bond angles
are slightly larger than 90°. This means that the s atomic orbital is not exactly
untouched; a small contribution from the s orbital is required to form near-
perpendicular sigma bonds and a LP. The energy of these non-bonding LP states
lies between the electron-bonding and antibonding energy levels.
Selenium chains have unpaired and unsatisfied valence states (DBs) at their
ends. Unlike a-Si:H, this localized state is not uncompensated, i.e. these orbitals
are doubly occupied by electrons, so we have C− 1 , where the minus superscript
refers to the atomic net charge and the subscript denotes the coordination number.
where u(k, r) describes the periodicity of the crystal lattice, i.e. u(k, r) =
u(k, r + R), where R is the lattice translation vector, and the exponential term
represents a plane wave. The wave vectors k of the electrons are related to the
translational symmetry of the lattice. It follows that a reciprocal lattice related
to the unit cell parameters is established in k-space. This solid state physics tool
4.2 Electronic density of states 91
where θ ik denotes the kth sp3 hybrid orbitals associated with the ith atom. The
first term is a sum over the interaction in which the kth and lth wave functions
belong to the same atom, and the second term is a sum associated with the wave
functions with the same bonds. These are the strongest interactions; this is why
the Weaire–Thorpe model is a good starting approximation: V1 V2 = 13 for
silicon. It can be shown that there is no gap for the case of V1 V2 = 12 only
(i.e. the valence and conduction bands touch each other).
Another important law for amorphous semiconductors is Anderson’s theory
of localization (Anderson, 1958). A simple description of this law is that the
increase in the disordered potential causes electron localization, i.e. the wave
function becomes confined into a small volume. This characteristic difference
between the electronic structure of crystalline and non-crystalline solids plays
an essential role in the transport and optical properties, etc.
92 Electronic structure
The rapid growth in the number of computer facilities and faster computa-
tional speeds mean that there are many different numerical methods that can
be used to calculate the electronic structure of condensed matter. The transla-
tional symmetry of crystalline materials is beneficial because of the application
of periodic boundary conditions, therefore only a few atoms are needed in the
numerical calculations to obtain an accurate result. For amorphous materials,
a periodic boundary condition using so-called “super cells,” is also applicable,
although there must be a large number of randomly distributed atoms to produce
reasonable results. Here, the term “large number” means at least 100 atoms.
Another possible method is cluster calculation, in which a structural model,
again containing a large number of atoms, is required. At the surface of these
clusters, the DBs should be saturated, usually with hydrogen termination. A
drawback of hydrogen termination is that a large charge transfer occurs in Si–H
bonds. To resolve this problem, a longer Si–H bond length than the average
normal Si–H atomic distance must be substituted into the model structure.
Three basic methods have been developed and used for the calculation of elec-
tronic properties of amorphous semiconductors: tight-binding models, Hartree–
Fock methods, and density-functional theories. For the electronic DOS calcu-
lation, the Hartree–Fock approximation usually overestimates the energy gap,
while the density-functional methods underestimates it.
The main conclusions of a large number of computer calculations and theo-
retical work are displayed in Figure 4.4. In the DOS of a fourfold-coordinated
pure amorphous semiconductor, delocalized and localized regions (tail states)
can be found separately by the mobility edge. Such localized band tails do not
exist in crystalline counterparts. Following Fermi statistics, the lower-energy
4.3 Defects 93
Figure 4.5. Non-bonding LP states are in the middle of the gap in the electronic
DOS.
4.3 Defects
Classification of defects
One possible definition of a defect is something that prohibits or prevents per-
fection. Structures that are devoid of defects are considered ideal. Inclusion of
a structural defect thus refers to a configuration in which an atom, or group
of atoms, does not satisfy the structural rules belonging to the ideal refer-
ence state of the material. Real materials contain structural defects that can
dominate their physical and chemical behaviors. In covalently bonded semi-
conductors, defects usually create states inside the energy gap which control
the electronic and optical properties. Some structural defects can capture elec-
trons or holes. It is not easy to identify structural defects in diffraction mea-
surements. The key defects that occur in crystalline materials comprise the
following:
r point defects due to impurity atoms in either substitutional or interstitial sites;
r complexities associated with special correlations between different point
defects such as impurity–vacancy pairs or donor–acceptor pairs;
r one-dimensional (line) defects, including translational displacements of atoms
(dislocations) and rotational displacements (disclinations);
94 Electronic structure
Figure 4.6. Defect model in a-Si. D0 (D+ ) and D− indicate neutral (positively
charged) and negatively charged dangling bonds. D+ and D− have no spin.
2C0 → C+ + C− . (4.5)
This means that the charged states C+ and C− are more stable than the neutral
state C0 . Kastner, Adler, and Fritzsche (1976) showed that these are threefold
(C+ −
3 ) and onefold (C1 ) coordinated in terms of the chemical bonds, and are
called the valence-alternation pairs (VAPs). Furthermore, another defect, the
threefold-coordinated Se atom (junction), can be formed in the network.
96 Electronic structure
Figure 4.7. Charge distribution of a 162 Se atom cluster calculated using the
DFT GGA method. Larger charge accumulations can be observed in both direc-
tions (see insets). (From Lukács, Hegedus, and Kugler (2009). J. Mater. Sci.:
Mater. Electron., 20, S33–S37. With kind permission from Springer Science
and Business Media.)
cannot be doped! Some years later, Spear and LeComber (1975) showed that
plasma-glow-discharge deposition of silicon, using mixtures of silane and
either phosphine or diborane, enabled electrically active pentavalent (P) (see
Figure 4.9) and trivalent (B) (see Figure 4.10) impurities to be incorporated
into the films, making them n type and p type, respectively. These impurities
introduce unsatisfied bonds into the network that are electronically active. An
interesting historical fact is that the first reported, but not successfully exploited,
experiment on doping had been carried out earlier by Chittick, Alexander, and
98 Electronic structure
Density of states
1 2 3 4
Figure 4.11. Electronic DOS obtained from Hartree–Fock ab initio cluster cal-
culations for different ta-C models doped by two nitrogen atoms. (From Kádas,
Ferenczy, and Kugler (1998). J. Non-Cryst. Solids, 227–230, 367. Copyright
2013 with permission from Elsevier.)
display the highest occupied molecular energies when two nitrogen atoms are
at the first- and fourth-neighbor positions, respectively. Peak 3 belongs to a ta-C
cluster containing only one nitrogen dopant. This latter configuration occurs
when the other dopant is placed at an infinite distance. These direct calculations
of the energy levels suggest that a significant difference can occur between
energy levels as a function of doping rate.
The diamond crystal structure consists of sixfold atomic rings, whereas the
corresponding amorphous diamond-like structure is made of five-, six-, and
sevenfold rings, with the threefold and fourfold rings being atypical. Never-
theless, triangle and squares break the local order, i.e. they can be considered
as one type of defect. Such defects were found in a-Si (Kugler et al., 2003).
As a consequence, two characteristic peaks can be found inside the tail of the
band structure (Kugler, 2012). The first larger peak corresponds to the square
atomic arrangements, whereas the peak at the higher energy is due to triangular
arrangements (Figure 4.12). In earlier investigations of electron transport, hop-
ping conductivity, optical properties, etc., the tail was usually considered to be
an exponential or Gaussian decaying function. Triangles and/or squares have
never been considered in any band structure calculations, although they play an
important role in several phenomena.
Normally a two-component (A and B) covalently bonded amorphous net-
work contains A–A, A–B, and B–B bonds. In special stoichiometric compos-
itions (GeSe2 or As2 Se3 , etc.) only A–B bonds are energetically favorable.
In these chalcogenide materials, a few A–A and B–B bonds can be found
100 Electronic structure
Density of states
Squares
Triangles
Energy
Defect spectroscopy
Defects produce localized states in the energy gap, and hence they can be
detected by different spectroscopic techniques. Since the defects dominate the
electronic and optical properties, it is important to estimate the number of
defects and their energy levels experimentally. Optical, electrical, and magnetic
measurements are the principal techniques used in defect spectroscopy, which
will be briefly reviewed in this section. The details of defect spectroscopy in
a-Si:H and amorphous chalcogenides (a-Chs) are described elsewhere (Singh
and Shimakawa, 2003).
Optical measurements
The optical absorption (the so-called “midgap absorption”) related to the defect
is usually weak (α 10 cm−1 ) in a-Si:H and a-Chs, and hence several sensitive
techniques are adopted, as described in the following.
of the pump beam is varied, the deflection of the probe beam becomes
a measure of the optical absorption spectrum of the materials. Some cal-
culations, however, are required to obtain the absolute optical absorption
coefficient. The PDS sensitivity is very high; for example, for a 1000 nm
(1 µm) thickness, values of α as low as 0.1 cm−1 can be obtained.
Electrical measurements
The recombination and emission of carriers at defects are investigated by elec-
tronic transport. As the thermal emission rate is related to the energy level, and
the shift of the Fermi level is related to the number of trapped carriers, these two
parameters are estimated in principle from the electrical transport measurements.
Several techniques are described in the following.
(i) Field-effect measurements were first applied to obtain the localized den-
sity of states in a-Si:H (Madan, LeComber, and Spear, 1976) and in
a-Chs (Marshall and Owen, 1976). A voltage is applied across the dielectric
thin layer deposited on the materials that yield band bending, producing a
space charge near the interface. Note that this is the same principle as for
a thin-film transistor. This produces an excess of carriers in the extended
states (excess conduction), which is due to the shift of the Fermi level
toward the band edge. The DOS near the Fermi level is estimated in this
method. As most of the space charges are 10 nm from the interface, they
are strongly affected by the presence of interface states, which may be a
drawback of this technique.
(ii) Capacitance measurements may be made by considering the depletion layer
of a Schottky barrier under reverse bias. As described in standard text
books, capacitance varies with the bias voltage, reflecting the space-charge
density, and the space charge (i.e. the localized carriers residing at the
defects) can be estimated from the capacitance value. The most familiar
technique that is used to measure deep levels in semiconductors is deep-
level transient spectroscopy (DLTS) in which the transient capacitance of
a Schottky barrier is measured (Lang, 1974). In DLTS, all the traps are
filled by applying a forward bias to the Schottky barrier, and subsequently a
reverse bias. The depletion layer width decreases over time to its steady state
value. Note that the capacitance is inversely proportional to the thickness
of the depletion layer.
4.3 Defects 103
Magnetic resonance
A quantum state, for example a defect state, occupied by a single electron is
split by a magnetic field; this is known as Zeeman splitting.
for a conventional magnetic field (3 T) between two Zeeman levels with
unpaired electrons whose spin states S = 12 (i.e. the paramagnetic state).
ESR signals can be observed by probing the microwave absorption if the
defects have unpaired spins larger than 1015 cm−3 , which of course depends
on the sensitivity of the equipment. A DB in a-Si:H has unpaired spin, and
in the so-called “device-quality” a-Si:H film the spin density is less than
1016 cm−3 (Street, 1991). When an unpaired electron interacts with nuclear
spin, the hyperfine structures are observed in the ESR spectra. Thus the
ESR spectra contain the information of the local configuration of these spin
states.
ESR signals in pure a-Chs are generally not observed under dark condi-
tions, because, unlike for a-Si:H, only the charged defect states (no unpaired
spins) are expected to exist in a-Chs (except for some components). Under
photoillumination at lower temperatures, however, ESR signals appear; we
call this light-induced ESR (LESR) (Bishop, Strom, and Taylor, 1975). Note
that LESR signals appear in a-Si:H at low temperature, which is not the case
for the original ESR signal under dark conditions (Street, 1991). Various
photoinduced effects have been monitored in experiments using LESR (Shi-
makawa, Kolobov, and Elliott, 1995). A pulsed ESR technique, in which
transient signals produced by coherent excitation using microwave pulses
are detected in the time domain, is also useful (Isoya et al., 1993).
(ii) Electron nuclear double resonance (ENDOR) is helpful in obtaining struc-
tural information about defects and their surroundings when hyperfine inter-
actions are weak. Measurements using ENDOR monitor the variation of
the ESR with nuclear magnetic resonance through hyperfine interaction.
The intensity of the ESR signal changes when NMR occurs. The details of
ENDOR measurements on a-Si:H have been described elsewhere (Morigaki,
1999).
where α = (Dc − 2)2 and β = (Dv − 2)2. Here Dc and Dv are the fractal dimen-
sionalities of the conduction and valence bands, respectively. Then eqn. (4.8)
produces (Nessa et al., 2000)
where 1n = α + β + 1. If the form of both Nc (E) and Nv (E) is parabolic, i.e.
α = β = 12 for 3D materials as in the most of the studies, then eqn. (4.11)
becomes
the tail (Kugler, 2012). The larger peak results from the squares, and the peak
with larger energy is formed by triangles.
C.B.
Epc
E
EF
V.B.
Figure 4.14. Potential fluctuation model in a-Si:H (Overhof and Thomas, 1989;
Shimakawa and Ganjoo, 2002).
induced in inhomogeneous media (Dyre, 1988; Dyre and Schroder, 2000). Both
the rf and THz ranges will be discussed in the following.
Usually, frequency-dependent conductivity (ac conductivity or ac loss) is
considered to be caused by a localized motion of carriers (Mott and Davis, 1979;
Singh and Shimakawa, 2003). A similar ac loss is also induced in macroscop-
ically inhomogeneous media (Dyre and Schroder, 2000; Singh and Shimakawa,
2003). In the following, we briefly summarize the ac loss in inhomogeneous
media, as this mechanism may dominate ac losses of a-Si:H in all frequency
ranges.
A very simple percolation-path method proposed by Dyre (1988) was origin-
ally applied to the random walk (hopping) of localized carriers; now it is also
applied to macroscopically inhomogeneous media. In inhomogeneous media,
“space charges” relax with a relaxation time τ = ɛ0 ɛ σ , where ɛ0 is the electric
constant, ɛ is the background dielectric constant, and σ is the electrical con-
ductivity. A space-charge relaxation induces the complex frequency-dependent
conductivity as (Dyre, 1988)
Figure 4.15. Real part of THz conductivity. The open circles and the solid
line show the experimental data and the theoretical prediction, respectively.
(From Shimakawa, Wagner, and Frumar (2013). Phys. Status Solidi B, 250,
1004–1007. Copyright 2013 by John Wiley and Sons Inc.)
inhomogeneity, are suggested in a-Si:H (Overhof and Thomas, 1989). The local
electronic conductivity therefore takes a variety of random values. Using
ln (a + ib) = ln a 2 + b2 + i tan−1 (b/a) (4.14)
τ max = 1.0 × 10−11 s and τ min = 4.3 × 10−13 s. The fitting of the predicted
data to the experimental data is good. Note that σ (0) = 0.35 S cm−1 does not
refer to the actual dc conductivity. Rather, it can be regarded as the virtual dc
conductivity, since, in the optical pumping technique, dc conductivity is not a
meaningful concept.
We now know that the value of τ min lies in the region of 10−13 s. Note again
that τ min is not deduced from the rf-range spectroscopy. Let us discuss what
τ min means. The ac loss in the rf range is interpreted under the assumption
of potential fluctuations (Shimakawa and Ganjoo, 2002). Here, the saturated
value of σ (ω) approaches 1 S cm−1 , and then the minimum dielectric relaxation
time τ min is expected to be 9 × 10−13 s. This value is close to that (4.3 × 10−13 s)
deduced from experimental data. Thus, it is suggested that potential fluctuations
which affect the rf conductivity also dominate conduction loss in the THz range
(Shimakawa, Wagner, and Frumar, 2013).
An alternative cause of THz conductivity – carrier hopping – which may
induce behavior similar to that previously described, should also be discussed.
The inverse of τ min (2.5 × 1012 s−1 ) is close to the phonon frequency, and Fekete
et al. (2009) have discussed this behavior in terms of a hopping-type relaxation
in localized tail states. It is not easy to distinguish which mechanism dominates
the THz conductivity: dielectric relaxation or hopping. If τ min changes with
temperature or optical pumping intensity, the dynamic loss may originate from
dielectric relaxation. There is unfortunately no literature available on this subject.
Photoluminescence
We briefly review the photoluminescence in a-Si:H and amorphous chalco-
genides. A radiative photon emission process in condensed matter is called
luminescence, and conceptually photoluminescence (PL) is the inverse pro-
cess of the optical absorption already discussed. As all information on photon
absorption and emission in materials can be involved in PL processes, PL is an
important topic, and there are many reports and reviews discussing PL mech-
anisms in amorphous semiconductors (see, for example, Aoki, 2006a, 2006b,
2012).
Frequency-resolved spectroscopy (FRS) may be used to understand directly
the lifetime distribution. In particular, quadrature frequency-resolved spec-
troscopy (QFRS) is known to be more suitable for studying amorphous semi-
conductors having a broad PL lifetime distribution (Aoki, 2006b). As shown
4.4 Optical and electronic properties 111
in Figure 4.16, Aoki and co-workers (Aoki, 2006a) succeeded in using QFRS
to expand the very wide time (frequency) range 10−9 –102 s in a-Si:H. There
are three main peaks in the lifetime distribution. The shortest lifetime (around
10−6 s) and the medium lifetime (around 10−3 s) can be identified as singlet (S)
and triplet (T) excitonic recombination. The longest lifetime (around 101 s) can
be explained by distant-pair recombination, in which photoexcited electrons and
holes are randomly distributed in band tails.
Similar S and T peaks have been also found in a-Chs (Aoki et al., 2005),
and the principal features in PL have been universally observed in amorphous
semiconductors.
107
50
As–Se–S
40 As–Se–I
5
10
As–Se–Te
In(0 S cm–1)
30
0 (S cm–1)
103 20
10
101 0
–10
10–1
0.0 0.2 0.4 0.6 0.8 1.0 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
Δ (eV) ΔE (eV)
(a) (b)
Figure 4.17. MN rule in (a) a-Si:H (Overhof and Thomas, 1989), (b) chalco-
genide glasses (Shimakawa and Abdel-Wahab, 1997).
References
Abkowitz, M. and Markovics, J.M. (1984). Evidence of equilibrium native defect popu-
lations in amorphous chalcogenides from analyses of xerographic spectra. Philos.
Mag. B, 49, L31–L36.
Anderson, P.W. (1958). Absence of diffusion in certain random lattices. Phys. Rev., 109,
1492–1505.
Anderson, P.W. (1975). Model for the electronic structure of amorphous semiconductors.
Phys. Rev. Lett., 34, 953–955.
Aoki, T. (2006a). Understanding the photoluminescence over 13-decade lifetime distri-
bution in a-Si:H. J. Non-Cryst. Solids, 352, 1138–1143.
Aoki, T. (2006b). Photoluminescence. In Optical Properties of Condensed Matter and
Applications, ed. J. Singh. Chichester: John Wiley & Sons, p. 75.
Aoki, T. (2012). Photoluminescence spectroscopy. In Characterization of Materials,
2nd edn., ed. E.N. Kaufmann. Hoboken, NJ: John Wiley & Sons, pp. 1–12.
Aoki, T., Saito, D., Ikeda, K., Kobayashi, S., and Shimakawa, K. (2005). Radiative
recombination processes in chalcogenide glasses deduced by lifetime measure-
ments over 11 decades. J. Opt. Adv. Mater., 7, 1749–1757.
Bishop, S.G., Strom, U., and Taylor, P.C. (1975). Optically induced localized paramag-
netic states in chalcogenide glasses. Phys. Rev. Lett., 34, 1346–1350.
Chittik, R.C., Alexander, J.H., and Sterling, H.J. (1969). The preparation and properties
of amorphous silicon. J. Electrochem. Soc., 116, 77–81.
Dyre, J.C. (1988). The random free energy barrier model for ac conduction in disordered
solids. J. Appl. Phys., 64, 2456–2468.
Dyre, J.C. and Schroder, T.B. (2000). Universality of ac conduction disordered solids.
Rev. Mod. Phys., 72, 873–892.
Fekete, L., Kuzel, P., Nemec, H. et al. (2009). Ultrafast carrier dynamics in microcrys-
talline silicon probed by time-resolved terahertz spectroscopy. Phys. Rev. B, 79,
115306, 1–13.
Ganjoo, A. and Shimakawa, K. (1994). Estimation of density of charged defects in
amorphous chalcogenides from a.c. conductivity: random-walk approach for bipo-
larons based on correlated barrier hopping. Philos. Mag. Lett., 70, 287–291.
114 Electronic structure
He, X.-F. (1990). Fractional dimensionality and fractional derivative spectra of interband
optical transitions. Phys. Rev. B, 42, 11751–11756.
Hirabayashi, I. and Morigaki, K. (1983). Light induced metastable effect on the short
lived photoinduced midgap absorption in hydrogenated amorphous silicon. J. Non-
Cryst. Solids, 59 & 60, 433–436.
Imagawa, O., Akiyama, T., and Shimakawa, K. (1984). Localized density of states in
amorphous silicon determined by electrophotography. Appl. Phys. Lett., 45, 438–
439.
Isoya, J., Yamasaki, S., Ohkushi, H., Matsuda, A., and Tanaka, K. (1993). Electron-
spin-echo envelope-modulation study of the distance between dangling bonds
and hydrogen atoms in hydrogenated amorphous silicon. Phys. Rev. B, 47, 7013–
7024.
Jackson, W.B. and Amer, N.M. (1982). Direct measurement of gap-state absorption in
hydrogenated amorphous silicon by photothermal deflection spectroscopy. Phys.
Rev. B, 25, 5559–5562.
Kádas, K. and Kugler, S. (1997a). Midgap states in nitrogen doped diamond-like amorph-
ous carbon. Solid State Commun., 102, 721–723.
Kádas, K. and Kugler, S. (1997b). Impurity levels in phosphorus and boron doped
amorphous silicon. Phil. Mag. B, 76, 281–285.
Kádas, K., Ferenczy, Gy.G., and Kugler, S. (1998). Theory of dopant pairs in four-
fold coordinated amorphous semiconductors. J. Non-Cryst. Solids, 227–230, 367–
371.
Kastner, M., Adler, D., and Fritzsche, H. (1976). Valence-alternation model for localized
gap states in lone-pair semiconductors. Phys. Rev. Lett., 37, 1504–1507.
Kugler, S. (2012). Advances in understanding the defects contributing to the tail states
in pure amorphous silicon. J. Non-Cryst. Solids, 358, 2060–2062.
Kugler, S. and Laszlo, I. (1989). Connection between topology and π -electron structure
in amorphous carbon. Phys. Rev. B, 39, 3882–3884.
Kugler, S., Kohary, K., Kádas, K., and Pusztai, L. (2003). Unusual atomic arrangements
in amorphous silicon. Solid State Commun., 127, 305–309.
Lang, D.V. (1974). Deep-level transient spectroscopy: a new method to characterize
traps in semiconductors. J. Appl. Phys., 45, 3023–3032.
Laszlo, I. (1993). Stable electronic energy levels in the presence of off-diagonal disorder.
Int. J. Quantum Chem., 48, 135–146.
Laszlo, I. (2000). Graph theoretical study of topologically determined electronic energy
levels. J. Mol. Struct.: THEOCHEM, 501–502, 501–508.
Linert, W. and Yelon, A. (eds.) (2013). Isokinetic relationships. Monatsh. Chem. (special
issue), 144, 1–2.
Lloyd-Hughes, J. and Tae-In Jeon (2012). A review of the terahertz conductivity of bulk
and nano-materials. J. Infrared Milli. Terahz. Waves, 33, 871–926.
Lukács, R., Hegedus, J., and Kugler, S. (2008). Structure and photoinduced volume
changes of obliquely deposited amorphous selenium. J. Appl. Phys., 104, 103512,
1–6.
Lukács, R., Hegedus, J., and Kugler, S. (2009). Microscopic and macroscopic models
of photo-induced volume changes in amorphous selenium. J. Mater. Sci: Mater.
Electron., 20, S33–S37.
Mackenzie, K.D., LeComber, L.G., and Spear, W.E. (1982). The density of states
in amorphous silicon determined by space-charge-limited current measurements.
Philos. Mag. B, 46, 377–389.
References 115
Madan, A., LeComber, P.G., and Spear, W.E. (1976). Investigation of the density of
localized states in a-Si using the field effect technique. J. Non-Cryst. Solids, 20,
239–257.
Mandelbrot, B.B. (1982). The Fractal Geometry of Nature. New York: Freeman.
Marshall, J.M. and Owen, A.E. (1976). Field-effect measurements in disordered
As30 Te48 Si12 Ge10 and As2 Te3 . Philos. Mag., 33, 457–474.
Morigaki, K. (1999). Physics of Amorphous Semiconductors. London: Imperial College
Press and World Scientific.
Mott, N.F. and Davis, E.A. (1979). Electronic Processes in Non-Crystalline Materials,
2nd edn. Oxford: Clarendon Press, p. 272.
Nessa, M., Shimakawa, K., Ganjoo, A., and Singh, J. (2000). Fundamental optical
absorption on fractals: a case example for amorphous chalcogenides. J. Optoelec-
tron. Adv. Mater., 2, 133–138.
Okamoto, H., Sobajima, Y., Toyama, T., and Matsuda, A. (2010). Laplace Meyer–Neldel
relation. Phys. Status Solidi A, 207, 566–569.
Overhof, H. and Thomas, P. (1989), Electronic Transport in Hydrogenated Amorphous
Semiconductors. Berlin: Springer-Verlag.
Pantelides, S.T. (1986). Defects in amorphous silicon: a new perspective. Phys. Rev.
Lett., 57, 2979–2982.
Petri, I., Salmon, P.S., and Fischer, H.E. (2000). Defects in a disordered world: the
structure of glassy GeSe2. Phys. Rev. Lett., 84, 2413–2416.
Rayment, T. and Elliott, S.R. (1983). Small-angle neutron scattering study of anisotropic
growth morphology and irreversible photodensification in a-GeSe3 films. Phys. Rev.
B, 28, 1174–1177.
Shimakawa, K. and Abdel-Wahab, F. (1997). The Meyer–Neldel rule in chalcogenide
glasses. Appl. Phys. Lett., 70, 652–654.
Shimakawa, K. and Aniya, M. (2013). Dynamics of atomic diffusion in condensed
matter: origin of the Meyer–Neldel compensation law. Monatsh. Chem., 144, 67–
71.
Shimakawa, K. and Ganjoo, A. (2002). ac photoconductivity of hydrogenated amorphous
silicon: influence of long-range potential fluctuations. Phys. Rev. B, 65, 165213,
1–5.
Shimakawa, K. and Katsuma, Y. (1986). Extended step-by-step analysis in space-charge-
limited current: application to hydrogenated amorphous silicon. J. Appl. Phys., 60,
1417–1421.
Shimakawa, K., Kolobov, A.V., and Elliott, S.R. (1995). Photoinduced effects and
metastability in amorphous semiconductors and insulators. Adv. Phys., 44, 475–
588.
Shimakawa, K., Singh, J., and O’Leary, S.K. (2006). Optical properties of disordered
condensed matter. In Optical Properties of Condensed Matter and Applications,
ed. J. Singh. Chichester: John Wiley & Sons, pp. 47–62.
Shimakawa, K., Wagner, T., and Frumar, M. (2013). THz photoconductivity in a-Si:H.
Phys. Status Solidi B, 250, 1004–1007.
Singh, J. (2002). Effective mass of charge carriers in amorphous semiconductors and its
applications. J. Non-Cryst. Solids, 299–302, 444–448.
Singh, J. and Shimakawa, K. (2003). Advances in Amorphous Semiconductors. London:
Taylor & Francis, p. 57.
Sólyom, J. (2009). Fundamentals of the Physics of Solids. Volume 2: Electronic Prop-
erties. Berlin Heidelberg: Springer-Verlag, p. 220.
116 Electronic structure
Spear, W.E. and Le Comber, P.G. (1975). Substitutional doping of amorphous silicon.
Solid State Commun., 17, 1193–1196.
Starbova, K., Mankov, V., Dikova, J., and Starbov, N. (1999). The effects of vapour
incidence on the microstructure and related properties of condensed GeS2 thin
films. Vacuum, 53, 441–446.
Street, R.A. (1991). Hydrogenated Amorphous Silicon. Cambridge: Cambridge Univer-
sity Press.
Tanaka, K. and Shimakawa, K. (2011). Amorphous Chalcogenide Semiconductors and
Related Materials. New York: Springer.
Vanecek, M., Kocka, J., Stuchlik, J., Kozisek, Z., Stika, O., and Triska, A. (1983). Density
of the gap states in undoped and doped glow discharge a-Si:H. Solar Energy Mater.,
8, 411–423.
Weaire, D. and Thorpe, M.F. (1971). Electronic properties of an amorphous solid. I.
A simple tight-binding theory. Phys. Rev. B, 4, 2508–2520.
Yelon, A., Movaghar, B., and Crandall, R.S. (2006). Multi-excitation entropy: its role in
thermodynamics and kinetics. Rep. Prog. Phys., 69, 1145–1194.
Zallen, R. (1983). The Physics of Amorphous Solids. New York: John Wiley & Sons Inc.
5
Photoinduced phenomena
117
118 Photoinduced phenomena
As most of the measurements of reversible PVE have been made after switch-
ing off the illumination (i.e. in the metastable state) (Hamanaka et al., 1976;
Tanaka, 1990, 1998; Kuzukawa et al., 1999), the correlation during illumination
remains unclear. To understand the dynamics of PVE and PD during illumin-
ation, in situ measurements were performed (Ganjoo, Ikeda, and Shimakawa,
1999; Ganjoo et al., 2000: Ganjoo and Shimakawa, 2002). A real-time in situ
surface height measuring system has been developed using a Twyman–Green
interferometer with image analysis technology (Ikeda and Shimakawa, 2004;
Shimakawa, Ikeda, and Kugler, 2004a).
Figure 5.1(a) shows an example of a surface height map for well-annealed
flatly deposited a-As2 Se3 films (on a Si substrate). The measurements were
carried out at 300 K. Figures 5.1(b) and (c) show the time evolution of
the changes in flatly and obliquely deposited films. In flatly deposited films
(Figure 5.1(b)), the surface height increased by 10 nm ( dd = 2%) dur-
ing 200 s of laser illumination (wavelength of 532 nm and power density of
90 mW cm−2 ). After 600 s, the illumination was turned off. The surface height
started decreasing and settled in 200 s, at which point the height was 2 nm lower
than before the light was switched off. In the case of obliquely deposited films
(Figure 5.1(c)), the surface height decreased by 12 nm ( dd = 2.4%) in 3 ×
104 s. Similar results have been observed in other a-Chs.
As shown in Figure 5.2, a less significant PVE than that seen in a-As2 Se3 was
found in (flatly deposited) a-Se films. The height increased rapidly by 2.5 nm
( dd = 0.5%) with the illumination (wavelength of 532 nm in wavelength and
power density of 90 mW cm−2 ). As soon as the light was turned off, after 800 s
of illumination, the height decreased by 2 nm and then gradually decreased to
the original height in 200 s.
Let us briefly discuss the height variations. As shown in Figure 5.1, in
flatly deposited a-As2 Se3 films PVE was observed during and after the illu-
mination. The transient PVE (PVE occurs only during the illumination) must
be involved in the whole PVE process (Ganjoo et al., 2000). The remain-
ing increase in surface height that occurred after illumination is a so-called
“metastable” PVE (Shimakawa, Kolobov, and Elliott, 1995; Tanaka, 1998). In
obliquely and as-deposited a-As2 Se3 films, as shown in Figure 5.1(c), the sur-
face decreased with time when the light was turned on (volume contraction).
The PVE may have been taken over by photoinduced volume contraction (PVC),
and hence the height decreased with time. This volume contraction can be inter-
preted by considering void corruption, as many voids are involved in obliquely
5.1 Photoinduced volume changes 119
[mm]
–2.0
–1.0
+2.0 0.0
+1.0
+2.0
+1.0
0.0
–1.0
–2.0 [nm]
[mm] 10
(a) 0
12
Light off
10
Height (nm)
8
6
4
2 Light on
0
–2
0 5 × 102 103 2 × 103 2 × 103 3 × 103 3 × 103
Time (s)
(b)
2
0
–2 Light on
Height (nm)
–4
–6
–8
–10
–12
–14
–16
0 104 2 × 104 3 × 104 4 × 104 5 × 104
Time (s)
(c)
Figure 5.1. (a) Surface height map for flatly deposited a-As2 Se3 . (b) Time
evolution of surface height for flatly deposited a-As2 Se3 . (c) Time evolution
of surface height for obliquely deposited a-As2 Se3 . (Adapted from Ikeda and
Shimakawa (2004). J. Non-Cryst. Solids, 338–340, 539. Copyright 2013 with
permission from Elsevier.)
120 Photoinduced phenomena
Figure 5.2. Time evolution of surface height for flatly deposited a-Se.
y
Figure 5.3. Snapshot of a glassy selenium network. The sample can move in
the z direction. (Adapted from Lukács, Hegedus, and Kugler (2009). J. Mater.
Sci.: Mater. Electron., 20, S33. With kind permission from Springer Science
and Business Media.)
the photon absorption, the electron and the hole became separated in space
on a femtosecond time scale (Hegedus et al., 2005). The Coulomb interaction
between the electron and the hole was neglected. The influences due to the
excited electron and the hole were treated independently. At first, an extra elec-
tron was put into the LUMO (excited electron creation), then an electron in
the HOMO (hole creation) was annihilated. When an additional electron was
put in the LUMO, a bond-breaking event occurred. In the majority of cases, a
covalent bond between twofold- and threefold-coordinated atoms was broken;
however, bond breaking was sometimes observed between twofold- and twofold-
coordinated atoms. Our localization analysis revealed that the LUMOs are usu-
ally localized at such sites before the bond breaking. The change in bond length
alternates between shrinkage and elongation in the vicinity of the broken bond
due to bond breaking. If the electron in the LUMO is de-excited, then all the
bond lengths are restored to their original value. A characteristic time devel-
opment of a single photoinduced bond-breaking event is shown in Figure 5.4.
Before the excitation at 5 ps, the bond length was approximately 0.255 nm. Dur-
ing illumination, this bond length increased by 10–20%, then decreased to its
original value after the de-excitation of the electron at 15 ps. The volume change
of the a-Se slab followed the bond breaking, and it showed damped oscillations
122 Photoinduced phenomena
3.4 25.6
3.2
25.5
thickness (A)
°
2.8
25.4
2.6
2.4 25.3
2.2
2.0 25.2
0 5 10 15 20
time (ps)
Figure 5.4. Atomic distance separation of breaking bond (solid line) and thick-
ness of sample (dotted line) as a function of time. Arrows indicate that an
excited electron was created at 5 ps and annihilated at 15 ps. (Taken with per-
mission from Hegedus et al. (2005). Phys. Rev. Lett., 95, 206803. Copyright
2013 by the American Physical Society.)
21.65
Light off Light off
length of the sample (Å)
21.6
21.55
21.5
21.45
Light on Light on
21.4
3 4 5 6
time (ps)
Figure 5.6. Difference between the Raman spectra of the 10 min illuminated
a-Se sample and that of the 40 min relaxed a-Se sample. (Taken with permis-
sion from Lukács et al. (2010). J. Appl. Phys., 107, 073517. Copyright 2013,
American Institute of Physics.)
to the vibrational modes of the covalent bonds in a-Se. As the Raman intensity
varies in time due to the illumination, a large number of covalent bonds break;
these form again once the illumination is switched off. The difference between
the Raman spectra of the sample illuminated for 10 min and that for the relaxed
sample after 40 min is shown in Figure 5.6. The Raman spectra measurement
provides strong experimental evidence for the predicted photoinduced bond-
breaking process.
A very different behavior is observed during hole creation (Hegedus et al.,
2005). Interchain weak bonds (see Figure 3.23) are formed after creating a
hole, thereby causing the contraction of the sample, as observed in Figure 5.7.
This always occurs near the atoms where the HOMO is localized. Since the
HOMO is usually localized in the vicinity of a onefold-coordinated atom, the
interchain bond formation often takes place between a onefold-coordinated atom
and a twofold-coordinated atom. Sometimes, the formation of interchain bonds
between two twofold-coordinated atoms is also observed.
124 Photoinduced phenomena
3.6 25.7
thickness (A)
3.4 25.6
°
3.2 25.5
3.0 25.4
0 5 10 15 20
time (ps)
Figure 5.7. Atomic distance separation of interchain bond (solid line) and thick-
ness of sample (dotted line) as a function of time. (Taken with permission from
Hegedus et al. (2005). Phys. Rev. Lett., 95, 206803. Copyright 2013 by the
American Physical Society.)
dn (t)
= G − Cn(t)2 . (5.2)
dt
5.1 Photoinduced volume changes 125
(a)
0
3
(b)
0
0 100 200 300 400
time (s)
G √
d(t) = β tanh GC t . (5.3)
C
The factor tanh(x) is measurable after a lengthy illumination (tanh() = 1).
√
Only one fitting parameter remains ( GC). After the light has been turned off,
the rate equation reduces to
dn (t)
= −Cn(t)2 , (5.4)
dt
which has the solution
a
d(t) = , (5.5)
(aCt/β) + 1
where a is equal to d(t) when the illumination is switched off. The photoinduced
expansion of a-Se films was measured in situ using optoelectronic interference,
enhanced by image processing (Ikeda and Shimakawa, 2004). Figures 5.8(a)
and (b) show the measured time evolution and decay, respectively, of the surface
height; the best fits were obtained in the 0–400 s interval of illumination and
after switching off the light.
126 Photoinduced phenomena
The reversible part follows the previously described time development dur-
ing and after illumination. There is no volume change caused by irreversible
microscopic effects after switching off the light. During the illumination, the
irreversible part of expansion/shrinkage is governed by
dn(t)
= G(N0 − n(t)), (5.7)
dt
where G describes the excited electron–hole generation rate and N0 is the max-
imum number of sigma bonds due to the irreversible bond breaking. Rewriting
the mathematical equation, we obtain
d(dirr (t))
= Girr − Cirr dirr (t). (5.8)
dt
The solution is given by
Girr
dirr(t) = (1 − exp(−Cirr t)). (5.9)
Cirr
The best fits of photoinduced volume changes during and after illumination
for flatly deposited a-AsSe are displayed in Figure 5.9. Both expressions have
positive prefactors; i.e. during the illumination, the sample expands.
The obliquely deposited porous sample shrinks. We can describe this process
using the same mathematical expression, but now both expressions have nega-
tive prefactors (Lukács and Kugler, 2011). The best fit for obliquely deposited
a-AsSe is shown in Figure 5.10. Negative prefactors mean that both reversible
and irreversible bond breaking cause macroscopic shrinkage in porous materials
(Chapter 3).
We should make an additional remark about time-dependent behavior. Instead
of a Debye-type normal exponential relaxation, the stretched exponential
5.1 Photoinduced volume changes 127
10 10
change of thickness (nm)
8 8
6 6
4 4
2 2
0 0
0 100 200 300 400 500 1000 1500 2000
(a) times (b) times
–2
–4
Height (nm)
–6
–8
–10
–12
–14
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) × 104
Figure 5.10. Volume changes as a function of time of obliquely deposited
a-AsSe thin film. Symbols denote the measured volume changes during the
illumination, and the solid line is the best fit. (From Lukács and Kugler (2011).
Jpn. J. Appl. Phys., 50, 091401.)
functions in many cases are fitted to the measured data in disordered systems.
The mathematical expression is given by
! "
D (t) = A 1 − exp − (t /τ )β , (5.10)
128 Photoinduced phenomena
–2
–4
Height (nm)
–6
–8
–10
–12
–14
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) × 104
Figure 5.11. Volume changes of the same obliquely deposited a-AsSe thin film
as in Figure 5.10 fitted by a stretched exponential function. (From Lukács and
Kugler (2011). Jpn. J. Appl. Phys., 50, 091401.)
where β (<1), A, and τ are fitting parameters. Usually these curves fit quite well,
yet the physical explanation of this phenomenon still remains unclear. In this
case, the stretched exponential function also fits the exponential data quite well,
as shown in Figure 5.11. This example demonstrates that two different processes
(reversible and irreversible) can also be fitted by a stretched exponential function.
5 20
4
15
Nd (×1015 cm–3)
ΔV/V (×10–6)
3
10
2
ΔV/V 5
1
Nd
0 0
0 1 2 3 4 5 6 7
Light soaking time (hr)
Figure 5.12. Time evolution of PVE together with PDC illuminated by Ar-ion
laser (300 mW cm−2 ) in a-Si:H. Open circles: volume change; closed squares:
defect density. (Taken with permission from Gotoh et al. (1998). Appl. Phys.
Lett., 72, 2978. Copyright 2013, American Institute of Physics.)
The opposite effect, i.e. volume compaction, occurs under bandgap illumin-
ation in g-SiO2 . Although the details are not clear, a possible mechanism is
related to PDC (Uchino et al., 2002). A photon creates a defect-like E center,
which may trigger the volume contraction. It is of interest to compare this effect
with volume expansion in a-Si:H.
(1) Among many a-Chs, a-As2 S3 shows the most significant change in the
bandgap, Eg ( Eg 50 meV at around room temperature).
(2) Eg decreases with increasing temperature and is scaled with TTg , where
T is the temperature at which the illumination is made and Tg is the glass-
transition temperature (Tanaka, 1983).
(3) Under the application of a high static pressure over 1.5 GPa (15 kbar), Eg
is known to decrease in a-Se (Tanaka, 1990).
(4) Bandgap illumination is the most effective for PD to occur.
50 K
2
Δα (cm–1) × 104 OFF
300 K
ON
0
0 2000 4000 6000 8000
Time (s)
Figure 5.13. Time evolution of α for a-As2 Se3 films during various cycles of
excitation at 50 and 300 K. (From Ganjoo et al. (2002). J. Non-Cryst. Solids,
299–302, 917. Copyright 2013 with permission from Elsevier.)
consistent with the ESR results, which evidently suggests the presence of well-
separated DBs. It is now known that the photoenhanced diffusion of H atoms
plays a role in PDC (Street, 1991; Branz, 1999; Morigaki, 1999; Morigaki and
Hikita, 2007). The above models highlight the importance of hydrogen diffu-
sion (H-diffusion) in a Si network which is assisted by photocarrier recombina-
tion energies. We should also consider the primary event of optical excitation;
optical excitation itself may induce bond breaking. We do not have a proper
answer to the question “Is a PDC process dominated by H-diffusion?”. Consider,
for example, the amorphous chalcogenides. Although the structural network is
completely different from that of a-Si:H, PDC does occur in a-Chs. Therefore,
it is suggested that the diffusion of hydrogen atoms is a secondary effect of PDC
processes.
We now discuss PDC in a-Chs. Shimakawa et al. (1990, 1995) have assumed
the existence of strong carrier–phonon interactions and have concluded that
the primary event of photoexcitation is the creation of a self-trapped exciton
(STE) and then its non-radiative recombination, leading either to the creation
of metastable STE (or an intimate pair of charged defects) or to its return to
the ground state through radiative recombination. This process is similar to that
proposed in a-SiO2 (Shugler and Stefanovich, 1990; Itoh and Stoneham, 2001).
According to the STE model, the primary event is an optical excitation, not
a non-radiative recombination. The weakening of a bond due to STE creation
eventually leads to bond breaking. However, it should be noted that the creation
of metastable STE is the result, not the cause, of bond breaking (Shugler and
Stefanovich, 1990).
Let us discuss the model in detail. As already described in Section 4.3, it
is widely accepted that defects in a-Chs are negative-U by their nature, and
hence only the charged defects exist in this class of material (Mott and Davis,
1979). These defects are ESR inactive. However, optical excitation (a change in
electron or hole occupation) changes a-Chs to ESR-active states. This is known
as light-induced ESR (LESR).
A microscopic model for PDC, for example for a-As2 S(Se)3 , is shown in
Figure 5.14. Optical excitation produces P+ − + −
2 − C1 and P4 − C1 , which are
called an intimate pair (IP) of defects, where P and C refer to pnictogen and
chalcogen centers, respectively. The IPs created in this way can be regarded as the
optically created STE (Shimakawa et al., 1995; Song and Williams, 1996). The
subsequent bond-switching reactions can lead to a significant separation between
the charged defects, which are called random pairs (RPs). RPs are responsible
for the decrease in the electronic properties (for example, photoconductivity,
5.3 Defect creation 133
Figure 5.14. A microscopic model for PDC for a-As2 S(Se)3 . X is the ground
state, Y1 and Y2 are the intimate pairs (IPs), and Z1 and Z2 are random pairs
(RPs).
LESR, etc.) as well as the optical properties (Shimakawa et al., 1995). Note that
these IPs and RPs disappear after thermal annealing at low temperature (around
200 K) and at high temperature (near Tg ). The details are described elsewhere
(Singh and Shimakawa, 2003; Tanaka and Shimakawa, 2011).
Let us discuss the quantum efficiency of PDC. As already explained, the use
of a photocurrent measurement during the illumination is useful in defect spec-
troscopy. Now let us focus on the defect creation N(t) during the illumination.
It is assumed that the change in the photocurrent is dominated by PDC itself.
The photocurrent, under conditions of thermal equilibrium, together with the
exposure time can be expressed as follows (Shimakawa et al., 2004b):
C C Ip (0)
Ip (t) = = = , (5.11)
N0 + N(t) N0 (1 + N(t)/N0 ) 1 + N(t)/N0
where C is a constant and Ip (0) (= CN0 ) is the initial photocurrent; N(t)N0 can
be estimated from the ratio Ip (0)Ip (t), with N0 being the initial number of
defects.
134 Photoinduced phenomena
Figure 5.15. Quantum efficiency η of defect creation at 300 K for both a-As2 Se3
(solid symbols) and a-Si:H (open symbols).
The photocurrent Ip (t) is written empirically for both a-Chs and a-Si:H as
follows (Shimakawa et al., 1995, 2004b):
Is
Ip (t) = ! ", (5.12)
1 − γ exp − (t /τ )β
where Is is a constant current attained by prolonged illumination, γ = Ns
(N0 + Ns ), τ is the effective creation time, and β is the dispersion parameter
(β < 1.0). Here Ns is the saturated number of the defect density. A change in Nt
is then given by
! "
N (t) = Ns 1 − exp − (t /τ )β . (5.13)
where G is the rate of absorbed photons (cm−3 s−1 ), t is the illumination time,
and np (= Gt) is the total number of absorbed photons.
For comparison, Figure 5.15 shows the experimental data of η vs. np at 300 K
for both a-As2 Se3 (solid symbols) and a-Si:H (open symbols). The efficiency η
decreased with increasing np and was higher for a-As2 Se3 than for a-Si:H. This
5.4 In situ measurements 135
can be attributed to the high degree of flexibility of the atomic structure in a-Chs.
A reported QE value of η 10−8 is found for a-Si:H, and the estimation here
from the photocurrent spectroscopy gives almost the same value after prolonged
illumination. Note that at the initial stage of the illumination the QE is much
higher than that after prolonged illumination for both a-As2 Se3 and a-Si:H.
(a) (b)
Figure 5.17. Time evolution of the change in N(t)N0 , α(t), and surface height
h(t) for thicknesses of (a) d = 700 nm and (b) d = 200 nm in a-As2 Se3 . (From
Nakagawa et al. (2010). Phys. Status Solidi C, 7, 857. Copyright 2013 by John
Wiley and Sons Inc.)
where A is a constant (equal to the total change), and τ and β have the same
meanings as described following eqn. (5.13). The fitting of eqn. (5.15) to the
experimental data (thin solid curves in Figures 5.17(a) and (b)) produces the
important physical parameters τ and β, which are indicated in Table 5.1.
In eqn. (5.15), the function with β < 1.0 is called the stretched exponential
function (SEF). As indicated in Table 5.1, the dynamics of PDC (N(t)N0 ) are
clearly represented by a SEF, although the origin of the SEF is not clear. On the
other hand, the dynamics of PVE ( h) are given by the exponential function
(β = 1.0). However, note that for the dynamics of PD ( α) when d = 700 nm,
β is 0.95, which is a lower value than that for d = 200 nm. We know that
PDC and PVE always follow stretched exponential and exponential dynamics,
5.4 In situ measurements 137
d = 200 nm d = 700 nm
τ (s) β τ (s) β
N(t)N0 18 0.55 40 0.55
h 70 1.0 90 1.0
α 200 1.0 180 0.95
PDC β<1
PVE β=1
PD β<1, β=1
Time
Figure 5.18. Comparison of the time scales in the evolutions of PDC, PVE,
and PD, together with the parameter β. (From Nakagawa et al. (2010). Phys.
Status Solidi C, 7, 857. Copyright 2013 by John Wiley and Sons Inc.)
References
Branz, H.M. (1999). Hydrogen collision model: quantitative description of metastability
in amorphous silicon. Phys. Rev. B, 59, 5498–5512.
Brazhkin, V.V., Gavrilyuk, A.G., Lyapin, A.G., Timofeev, Yu.A., Katayama, Y., and
Kohara, S. (2007). AsS: bulk inorganic molecular-based chalcogenide glasses.
Appl. Phys. Lett., 91, 031912, 1–3.
Charnovych, S. and Kökényesi, S. (2011). Plasmon enhanced optical recording in As–Se
layers. Phys. Status Solidi C, 8, 2854–2857.
Charnovych, S., Kökényesi S., Glodan, Gy., and Csik, A. (2011). Enhancement of
photoinduced transformations in amorphous chalcogenide film via surface plasmon
resonances. Thin Solid Films, 519, 4309–4312.
Charnovych, S., Szabó, I.A., Tóth, A., Volk, J., Trunov, M.L., and Kökényesi, S. (2013).
Plasmon assisted photoinduced surface changes in amorphous chalcogenide layer.
J. Non-377Cryst. Solids, 200–204.
Cserhati, C., Charnovych, S., Lytvyn, P.M. et al. (2012). Mechanism of photo induced
mass transfer in amorphous chalcogenide films. Mater. Lett., 66, 159–161.
Dresner, J. and Stringfellow, G.B. (1968). Electronic processes in the photo-
crystallization of vitreous selenium. J. Phys. Chem. Solids, 29, 303–311.
Elliott, S.R. (1986). A unified model for reversible photostructural effects in chalco-
genide glasses. J. Non-Cryst. Solids, 81, 71–98.
Elliott, S.R. and Kolobov, A.V. (1991). Athermal light-induced vitrification of
As50 Se50 films. J. Non-Cryst. Solids, 128, 216–220.
142 Photoinduced phenomena
Freitas, R.J., Shimakawa, K., and Wagner, T. (2014). The dynamics of photoinduced
defect creation in amorphous chalcogenides: the origin of the stretched exponential
function. J. Appl. Phys., 115, 013704, 1–4.
Fritzshe, H. (2000). Light-induced effects in glasses. In Insulating and Semiconducting
Glasses, ed. P. Boolchand. Singapore: World Scientific, pp. 653–690.
Ganjoo, A. and Shimakawa, K. (2002). Dynamics of photodarkening in amorphous
chalcogenides. J. Optoelectron. Adv. Mater., 4, 595–604.
Ganjoo, A., Ikeda, Y., and Shimakawa, K. (1999). In situ photoexpansion measurements
of amorphous As2 S3 films: role of photocarriers. Appl. Phys. Lett., 74, 2119–2122.
Ganjoo, A., Shimakawa, K., Kamiya, H., Davis, E.A., and Singh, J. (2000). Percolative
growth of photodarkening in amorphous As2 S3 films. Phys. Rev. B, 62, R14601–
14604.
Ganjoo, A., Shimakawa, K., Kitano, K., and Davis, E.A. (2002). Transient photodark-
ening in amorphous chalcogenides. J. Non-Cryst. Solids, 299–302, 917–923.
Gotoh, T., Nonomura, S., Nishio, M., Nitta, S., Kondo, M., and Matsuda, A. (1998).
Experimental evidence of photoinduced expansion in hydrogenated amorphous
silicon using bending detected optical lever method. Appl. Phys. Lett., 72, 2978–
2980.
Hamanaka, H., Tanaka, K., Matsuda, A., and Iijima, S. (1976). Reversible photo-induced
changes in evaporated As2 S3 and As4 Se5 Ge1 films. Solid State Commun., 19, 499–
501.
Hegedus, J., Kohary, K., Pettifor, D.G., Shimakawa, K., and Kugler, S. (2005). Photo-
induced volume changes in amorphous selenium. Phys. Rev. Lett., 95, 206803, 1–4.
Hegedus, J., Kohary, K., and Kugler, S. (2006). Universal feature of photo-induced
volume changes in chalcogenide glasses. J. Non-Cryst. Solids, 352, 1587–1590.
Ikeda, Y. and Shimakawa, K. (2004). Real-time in situ measurements of photoinduced
volume changes in chalcogenide glasses. J. Non-Cryst. Solids, 338–340, 539–
542.
Inami, T. and Adachi, S. (1999). Structural and optical properties of photocrystallized
Se films. Phys. Rev. B, 60, 8284–8289.
Itoh, N. and Stoneham, A.M. (2001). Materials Modification by Electronic Excitation.
Cambridge: Cambridge University Press.
Kaganovskii, Yu., Beke, D.L., Charnovych, S., Kökényesi, S., and Trunov, M.L. (2011)
Inversion of the direction of photo-induced mass transport in As20 Se80 films: ex-
periment and theory. J. Appl. Phys., 110, 063502, 1–5.
Kökényesi, S. (2006). Amorphous chalcogenide nano-multilayers: research and devel-
opment. J. Optoelectron. Adv. Mater., 8, 2093–2096.
Kolobov, A.V. (ed.) (2003). Photo-Induced Metastability in Amorphous Semiconductors.
Weinheim: Wiley-VCH.
Kolobov, A.V. and Elliott, S.R. (1995). Reversible photo-amorphization of a crystallized
As50 Se50 alloy. Philos. Mag. B, 71, 1–10.
Kolobov, A.V., Oyanagi, H., Tanaka, K., and Tanaka, K. (1997). Structural study of
amorphous selenium by in-situ EXAFS: observation of photoinduced bond alter-
nation. Phys. Rev. B, 55, 726–734.
Kugler, S., Hegedus, J., and Kohary, K. (2007). Modelling of photoinduced changes
in chalcogenide glasses: a-Se and a-As2 Se3 . J. Mater. Sci.: Mater. Electron., 18,
S163–S167.
Kuzukawa, Y., Ganjoo, A., Shimakawa, K., and Ikeda, Y. (1999). Photo-induced struc-
tural changes in obliquely deposited arsenic-based amorphous chalcogenides: a
model for photostructural changes. Philos. Mag. B, 79, 249–256.
References 143
Lukács, R. and Kugler, S. (2011). Photoinduced volume changes of obliquely and flatly
deposited amorphous AsSe films: universal description of the kinetics. Jpn. J. Appl.
Phys., 50, 091401, 1–4.
Lukács, R., Hegedus, S., and Kugler, S. (2009). Microscopic and macroscopic models
of photo-induced volume changes in amorphous selenium. J. Mater. Sci.: Mater.
Electron., 20, S33–S37.
Lukács, R., Veres, M., Shimakawa, K., and Kugler, S. (2010). On photoinduced vol-
ume change in amorphous selenium: quantum chemical calculation and Raman
spectroscopy. J. Appl. Phys., 107, 073517, 1–5.
Mikla, V.I. and Mikhalko, I.P. (1995). Laser-induced structural transformation of
Asx Se1−x thin amorphous films. J. Non-Cryst. Solids, 180, 236–243.
Molina, D., Lomba, E., and Kahl, G. (1999). Tight-binding model of selenium disordered
phases. Phys. Rev. B, 60, 6372–6382.
Morigaki, K. (1999). Physics of Amorphous Semiconductors. London: World Scientific
& Imperial College Press.
Morigaki, K. and Hikita, H. (2007). Modeling of light-induced defect creation in hydro-
genated amorphous silicon. Phys. Rev. B, 76, 085201, 1–17.
Mott, N.F. and Davis, E.A. (1979). Electronic Processes in Non-Crystalline Materials,
2nd edn. Oxford: Oxford University Press.
Naik, R., Adarsh, K.V., Ganesan, R. et al. (2009). X-ray photoelectron spectroscopic
studies on Se/As2 S3 and Sb/As2 S3 nanomultilayered film. J. Non-Cryst. Solids,
355, 1836–1839.
Nakagawa, N., Shimakawa, K., Itoh, T., and Ikeda, Y. (2010). Dynamics of principal
photoinduced effects in amorphous chalcogenides: in situ simultaneous measure-
ments of photodarkening, volume changes, and defect creation. Phys. Status Solidi
C, 7, 857–860.
Palmer, R.G., Stein, D.L., Abraham, E., and Anderson, P.W. (1984). Models of hier-
archically constrained dynamics for glassy relaxation. Phys. Rev. Lett., 53, 958–
961.
Pfeifer, G., Paesler, M.A., and Agarwal, S.C. (1991). Reversible photodarkening of
amorphous arsenic chalcogens. J. Non-Cryst. Solids, 130, 111–143.
Phillips, J.C. (1996). Stretched exponential relaxation in molecular and electronic
glasses. Rep. Prog. Phys., 59, 1133–1207.
Redfield, D. and Bube, R.H. (1996). Photoinduced Defects in Semiconductors. New
York: Cambridge University Press.
Sakai, K., Maeda, K., Yokoyama, H., and Ikari, T. (2003). Photo-enhanced crystallization
by laser irradiation and thermal annealing in amorphous GeSe2 . J. Non-Cryst.
Solids, 320, 223–230.
Shimakawa, K. (1984). Origin of nonsymmetric dielectric relaxation in dipolar materials.
Appl. Phys. Lett., 45, 587–588.
Shimakawa, K. (1985). Exciton recombination in amorphous chalcogenides. Phys. Rev.
B, 31, 4012–4014.
Shimakawa, K. (1986). Persistent photocurrent in amorphous chalcogenides. Phys. Rev.
B, 34, 8703–8708.
Shimakawa, K., Inami, S., and Elliott, S.R. (1990). Reversible photoinduced change
of photoconductivity in amorphous chalcogenide films. Phys. Rev. B, 42, 11857–
11861.
Shimakawa, K., Inami, S., Kato, T., and Elliott, S.R. (1992). Origin of photoin-
duced metastable defects in amorphous chalcogenides. Phys. Rev. B, 46, 10062–
10069.
144 Photoinduced phenomena
Shimakawa, K., Kolobov, A.V., and Elliott, S.R. (1995). Photoinduced effects and
metastability in amorphous semiconductors and insulators. Adv. Phys., 44, 475–
588.
Shimakawa, K., Yoshida, N., Ganjoo, A., Kuzukawa, Y., and Singh, J. (1998). A model
for the photostructural changes in amorphous chalcogenides. Philos. Mag. Lett.,
77, 153–158.
Shimakawa, K., Ikeda, Y., and Kugler, S. (2004a). Fundamental optoelectronic processes
in amorphous chalcogenides. In Optoelectronic Materials and Devices, Vol.1, eds.
G. Lucovsky and M. Popescu. Bucharest: INOE Publishing House, p. 103.
Shimakawa, K., Mehern-Nessa, Ishida, H., and Ganjoo, A. (2004b). Quantum efficiency
of light-induced defect creation in hydrogenated amorphous silicon and amorphous
As2 Se3 . Philos. Mag. B, 84, 81–89.
Shimakawa, K., Nakagawa, N., and Itoh, T. (2009). The origin of stretched exponential
function in dynamic response of photodarkening in amorphous chalcogenides.
Appl. Phys. Lett., 95, 051908, 1–3.
Shluger, A. and Stefanovich, E. (1990). Models of the self-trapped exciton and nearest-
neighbor defect pair in SiO2 . Phys. Rev. B, 42, 9664–9673.
Singh, J. and Shimakawa, K. (2003). Advances in Amorphous Semiconductors. New
York and London: Taylor and Francis.
Song, K.S. and Williams, R.T. (1996). Self-Trapped Excitons, 2nd edn. Berlin: Springer.
Staebler, D.L. and Wronski, C.R. (1977). Reversible conductivity changes in discharge-
produced amorphous Si. Appl. Phys. Lett., 31, 292–294.
Street, R.A. (1991). Hydrogenated Amorphous Silicon. Cambridge: Cambridge Univer-
sity Press.
Takats, V., Miller, F., Jain, H., Cserhati, C., and Kökényesi, S. (2009). Direct surface
patterning of homogeneous and nanostructured chalcogenide layers. Phys. Status
Solidi C, 6(suppl. 1), S83–S85.
Tanaka, K. (1983). Mechanisms of photodarkening in amorphous chalcogenides.
J. Non-Cryst. Solids, 59–60, 925–928.
Tanaka, K. (1990). Photoinduced structural changes in chalcogenide glasses. Rev. Solid
State Sci., 4, 641–659.
Tanaka, K. (1998). Photoexpansion in As2 S3 glass. Phys. Rev. B, 57, 5163–5167.
Tanaka, K. and Shimakawa, K. (2011). Amorphous Chalcogenide Semiconductors and
Related Materials. New York: Springer.
Trunov, M.L., Lytvyn, P.M., Yannopoulos, S.N., Szabo, I.A., and Kökényesi, S. (2011).
Photoinduced mass-transport based holographic recording of surface relief gratings
in amorphous selenium films. Appl. Phys. Lett., 99, 051906, 1–3.
Uchino, T., Takahashi, T., Ichii, M., and Yoko, T. (2002). Microscopic model of photoin-
duced and pressure induced UV spectra changes in germanosilicate glass. Phys.
Rev. B, 65, 172202, 1–4.
Voynarovych, I., Kökényesi, S., Yurkovich, N., Charnovych, S., and Dmitruk, N. (2012).
Plasmon-assisted transformations in metal-amorphous chalcogenide light-sensitive
nanostructures. Plasmonics, 7, 341–345.
Index
145
146 Index