Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Physics of Thin-Film Photovoltaics

Victor G. Karpov
Visit to download the full and correct content document:
https://ebookmass.com/product/physics-of-thin-film-photovoltaics-victor-g-karpov/
More products digital (pdf, epub, mobi) instant
download maybe you interests ...

Conceptual Physics Pearson International Edition Paul


G. Hewitt

https://ebookmass.com/product/conceptual-physics-pearson-
international-edition-paul-g-hewitt/

Microwave and Millimeter-Wave Chips Based on Thin-Film


Integrated Passive Device Technology. Design and
Simulation Yongle Wu

https://ebookmass.com/product/microwave-and-millimeter-wave-
chips-based-on-thin-film-integrated-passive-device-technology-
design-and-simulation-yongle-wu/

Thin Objects Øystein Linnebo

https://ebookmass.com/product/thin-objects-oystein-linnebo/

McEvoy's Handbook of Photovoltaics, Third Edition:


Fundamentals and Applications Soteris Kalogirou

https://ebookmass.com/product/mcevoys-handbook-of-photovoltaics-
third-edition-fundamentals-and-applications-soteris-kalogirou/
HTML_FOR_WEB_DEVELOPMENT Henry Victor

https://ebookmass.com/product/html_for_web_development-henry-
victor/

Monster Problems Victor Pineiro

https://ebookmass.com/product/monster-problems-victor-pineiro/

Monster Problems Victor Piñeiro

https://ebookmass.com/product/monster-problems-victor-pineiro-2/

The Maternal Imagination of Film and Film Theory 1st


ed. Edition Lauren Bliss

https://ebookmass.com/product/the-maternal-imagination-of-film-
and-film-theory-1st-ed-edition-lauren-bliss/

These Thin Lines 1st Edition Milena Mckay

https://ebookmass.com/product/these-thin-lines-1st-edition-
milena-mckay/
Physics of Thin-Film
Photovoltaics
Scrivener Publishing
100 Cummings Center, Suite 541J
Beverly, MA 01915-6106

Publishers at Scrivener
Martin Scrivener (martin@scrivenerpublishing.com)
Phillip Carmical (pcarmical@scrivenerpublishing.com)
Physics of Thin-Film
Photovoltaics

Victor Karpov
and
Diana Shvydka
This edition first published 2022 by John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
and Scrivener Publishing LLC, 100 Cummings Center, Suite 541J, Beverly, MA 01915, USA
© 2022 Scrivener Publishing LLC
For more information about Scrivener publications please visit www.scrivenerpublishing.com.

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or other-
wise, except as permitted by law. Advice on how to obtain permission to reuse material from this title
is available at http://www.wiley.com/go/permissions.

Wiley Global Headquarters


111 River Street, Hoboken, NJ 07030, USA

For details of our global editorial offices, customer services, and more information about Wiley prod-
ucts visit us at www.wiley.com.

Limit of Liability/Disclaimer of Warranty


While the publisher and authors have used their best efforts in preparing this work, they make no rep­
resentations or warranties with respect to the accuracy or completeness of the contents of this work and
specifically disclaim all warranties, including without limitation any implied warranties of merchant-­
ability or fitness for a particular purpose. No warranty may be created or extended by sales representa­
tives, written sales materials, or promotional statements for this work. The fact that an organization,
website, or product is referred to in this work as a citation and/or potential source of further informa­
tion does not mean that the publisher and authors endorse the information or services the organiza­
tion, website, or product may provide or recommendations it may make. This work is sold with the
understanding that the publisher is not engaged in rendering professional services. The advice and
strategies contained herein may not be suitable for your situation. You should consult with a specialist
where appropriate. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other damages.
Further, readers should be aware that websites listed in this work may have changed or disappeared
between when this work was written and when it is read.

Library of Congress Cataloging-in-Publication Data

ISBN 9781119651000

Cover image: Pixabay.com


Cover design by Russell Richardson

Set in size of 11pt and Minion Pro by Manila Typesetting Company, Makati, Philippines

Printed in the USA

10 9 8 7 6 5 4 3 2 1
Dedication

To our parents

v
Contents

Preface xi
Part I General and Thin Film PV 1
I. Introduction to Thin Film PV 1
A. The Origin of PV. Junctions 1
B. Fundamental Material Requirements 3
C. Charge Transport. Definition of Thin Film PV 4
D. Distinctive Features of Thin Film PV 7
References 11
Part II One-Dimensional (1D) Diodes and PV 13
II. 1D Diode 13
A. Metal-Insulator-Metal Diode 13
B. Schottky, Reach-Through, and
Field-Compensation Diodes19
1. Schottky Diode 19
2. Reach Through Diodes 21
3. Field Compensation Diode 23
C. P-N Homo-Junctions 24
D. Heterojunctions 26
E. Other Relevant Types of Diodes 28
F. Field Reversal Diode: A Counterintuitive Case 29
G. Cat’s Whisker Diode 30
III. 1D Solar Cell 32
A. 1D Solar Cell Base Model 32
B. Numerical Modeling of 1D PV 38
1. Governing Equations 39
2. Device Model Parameters 40
3. Some Modeling Results 42
IV. Photovoltaic Parameters 43
A. Second-Level Parameters 44
B. Practical Solar Cells and Third-Level Metrics 46

vii
viii Contents

C. Indicative Facts 49
D.  Phenomenological Interpretation. Ideal
Diode with Other Circuitry Elements 52
V. Case Study 54
A. Field Reversal PV 54
1. Analytical Approach 55
2. Numerical Modeling of the Field
Reversal Device Operations 60
B. Miraculous Back Contact 68
References 72
Part III Beyond 1D: Lateral Effects in Thin Film PV 79
VI. Examples of Multidimensional Numerical Modeling 79
VII. Introduction to Random Multidimensional
Phenomena81
VIII. Lateral Screening Length 84
A. Shunt Screening 84
B. Bias Screening 85
C. Quantitative Approach and Linear
Screening Regime 88
IX. Schottky Barrier Nonuniformities 91
X. Semi-Shunts 93
XI. Random Diodes 96
A. Weak Diodes 96
B. Random Diode Arrays in Solar Cells 99
C. Random Diode Arrays in PV Modules
and Fields 105
XII. Nonuniformity Observations 109
A. Cell Level Observations 109
B. Module Level Observations 118
XIII. Nonuniformity Treatment 121
References 125
Part IV Electronic Processes in Materials of Thin Film PV 131
XIV. Morphology, Fluctuations, and the Density of States 132
A.  The Materials of Thin Film PV
are Fundamentally Different 132
B. Noncrystalline Morphology 134
C. Long Range Fluctuations of Potential Energy 136
D. Random Potential in Very Thin Structures 139
E. Numerical Estimates and Implications 142
XV. Electronic Transport 144
Contents ix

A. Band Transport in Random Potential 144


B. Hopping Transport Through Thin
Noncrystalline Films 147
1. Hopping Between Ideal Electrodes 149
2. Hopping Between Resistive Electrodes 151
3. Critical Area and Mesoscopic Fluctuations 153
XVI. Recombination in Quasi-Continuous Spectrum 155
XVII. Noncrystalline Junctions 161
XVIII. Piezo and Pyro-PV 164
A. The Nature of Piezo-PV 164
B. Piezo-PV Observations 169
C. The Significance of Piezo-PV 171
References 174
Part V Electro-Thermal Instabilities in Thin Film PV 181
XIX. The Two-Diode Model 182
A. Linear Stability Analysis 183
B. The Two-Diode Modeling: Numerical
Estimates and Scaling 184
XX. Distributed Diode Model 186
A. Introduction 186
B. Linear Stability Analysis 187
XXI. Simplistic Numerical Modeling 188
XXII. Spontaneous Hot Spots 190
A. Introduction 190
B. Observations 191
C. Numerical Modeling 195
1. Electrical Model 195
2. Thermal Model 199
D. Modeling Results 200
E. Approximate Analytical Model 205
XXIII. Related Work 207
XXIV. Conclusions on the Electro-Thermal
Instabilities in Thin Film PV 209
References 210
Part VI Degradation of Thin Film PV 213
XXV. Thin Film vs Crystalline PV Degradation
Processes213
XXVI. Observations 215
A. Cell Degradation 216
B. Module Degradation 222
x Contents

XXVII. Categories of Degradation 225


A. General Categories 225
B. Thin-Film PV Instabilities 227
1. Shunting Instability 227
2. Contact Delamination Instability 229
XXVIII. Accelerated Life Testing 231
A. Examples of Very Strong ALT: HALT 232
1. EBIC HALT 232
2. LBIC HALT 234
B. Actuarial Approach to ALT 235
C. Concluding Remarks on Degradation 236
References 237
Appendix. Some Methodological Aspects of Device Modeling 243
Appendix A: Model of Series Connection 243
Appendix B: The Diffusion Approximation 245
Appendix C: Long Range Potential 248
1. Point Changes 248
2. Columnar Charges 251
References 253
Index 255
Preface

There is no longer a need to argue about the importance of solar energy and
the necessity of furthering the photovoltaic (PV) industry. These issues have
been addressed by many sources in the media and publications. Excellent
books have been published covering the basics of photovoltaics including
both the underlying classical physics and material implementations.
Taking advantage of the above issues sufficiently presented, this book
will concentrate on several subjects left beyond the scope of the exiting
photovoltaic texts. These subjects are all related to thin film photovoltaics
(such as CdTe, CIGS, or a-Si:H based) whose properties and operations
turn out to be quite different from that of the classical PV presented mostly
by the crystalline Si structures. The obvious differences lie in the device
thinness (microns instead of millimeters) and its morphology (polycrys-
talline or amorphous instead of crystalline).
The thinness effect may be so significant that the photogenerated charge
carriers reach the electrodes without much recombination even in the
imperfect non-crystalline material, which deemphasizes the classical con-
cept of recombination limited PV performance. On the other hand, the
transversal resistance not self-averaging across thin disordered structures
leaves a possibility of significant lateral nonuniformities, some of which
can be quite detrimental. In addition, the non-crystalline morphology
leads to continuous energy spectra of localized states instead of discrete
levels in crystals, which results in new transport mechanisms (hopping)
and recombination features. Finally, an important part of thin film PV
possess piezo-electric properties, which leads to the concept of piezo-PV
unknown in the classical PV science.
Taken together, the latter subjects form the physics of thin film PV as a
distinctive field of its own. This book will for the first time provide a con-
sistent presentation of that field. In other words, this book will not describe
the material of classical PV science, which has been masterfully described
in other PV books, but will instead introduce material that has never been
presented in PV books before.

xi
xii Preface

To better explain our book title, we would like to emphasize the term
of physics as defining the filed that is not necessarily related to the object’s
materials, chemistry or engineering. The physics was always related to phe-
nomena per se, regardless of the material specificity. Think, for example,
of the Newton’s second law where the material structure does not matter,
think about the fact that the Planck’s constant and Boltzmann’s constant
do not depend on which material they describe, and how, in the physics of
disordered systems, it does not matter what is the chemical composition
of a system, etc. Similarly, in the photovoltaic science, the basics (built-in
field, junction, and so on) remain material unspecific. Along these lines,
our book concentrates on the aspects that are not sensitive to the material
structure or composition. That is not to denigrate the known great contri-
butions to photovoltaics from the materials sciences and chemistry, which
appear dominant from any query of recent publications, but rather to pref-
ace our book limitations.
In our experience, the concept of a distinctive field of thin film PV phys-
ics may not appear obvious to everybody: quite a few in the community
believe that thin film PV must be understood in the framework of classical
PV science developed earlier and proven for thick crystalline systems. Such
a resistance appears rather paradoxical given that thin film PV is made of
materials that seem unacceptable from the classical PV perspective, and
yet they often outperform classical PV. Our book solves that paradox by
switching to a new physics paradigm.
The book is naturally broken into six parts, each containing several
interrelated sections.
Part I, consisting of just section I, gives a general introduction to PV
including the concepts of junctions, material requirements, and distinctive
features of thin film PV.
Part II, encompassing sections II to V, concentrates on one-dimensional
concepts relevant to thin film PV. It presents a densely populated zoo of
various diodes, some quite different from p-n junctions, including the ones
that play a significant role in thin film PV operations. It then extends that
diode consideration on one-dimensional analysis of solar cells, including
some important case studies.
Part III, made up of sections VI to XIII, takes the consideration beyond
one-dimensional physics considering lateral effects, such as shunting, ran-
dom micro-diodes, weak diodes, etc., addressing both the related theory
and observations.
Part IV, including sections XIV to XVIII, concentrates on electronic pro-
cesses in disordered materials of thin film PV which addresses morphology,
fluctuations, transport properties, recombination in the quasi-­continuous
Preface xiii

spectrum of localized states, operations of nonocrystalline junctions, and,


finally, piezo photovoltaics.
Part V, formed by sections XIX to XXIV, presents the subject of
­electro-thermal instabilities in thin film PV starting from the exactly
solved two-diode model and current hogging to spontaneously formed hot
spots conducive of performance degradation.
Part VI includes sections XXV to XXVIII and discusses a sensitive issue of
device degradation for thin film PV, including both solar cells and PV modules.
It then discusses the problematic of accelerated life testing and risk analyses.
Finally, our book has an Appendix that comprise three parts of method-
ical value” the validity of in-series representation of complex band dia-
grams, the nature of saturation current in diode models, and accurate
estimates for the electron potential fluctuations.
We are grateful to many people who have helped us to learn the subjects
presented in this book. Our exposure to PV science and industry started
with First Solar LLC (Perrysburg, OH) before the year 2000, at which time
its research and production groups combined many enthusiastic contrib-
utors including H. McMaster, G. Dorer, G. Nelson, R. C. Powell, D. Rose,
U. Jayamaha, E. Bykov, T. Maxon, R. Harju, T. Colman, L. McFaul, N. Reiter,
T. Kahle, G. Khouri, G. Rich, M. Steel, and many others, to all of whom we
are grateful for their intelligent friendly support. We would like as well
to acknowledge our interactions with the PV group of the University of
Toledo including A.D. Compaan, X. Deng, A. Vijh, D. Giolando, A. Vasko,
Y. Roussillon, V. Parikh, L. Attigalle (Cooray), M. Mitra, M. Nardone, X. Li,
V. Plotnikov, K. Wieland, and J. Drayton. Finally, we would like to express
our gratitude to the members of CdTe PV National Team at NREL, starting
with their remarkable managers K. Zweibel, B. von Roedern, H. Ulllal, and
brilliant contributors B. McCandless, D. Albin, V. Kaydanov, T. Ohno, J.
Sites, T. Sampath, A. Fahrenbruch, and so many others.
This book is intended for three readership categories. One is that of
graduate and advanced undergraduate students with some understanding
of the general physics and familiarity with the basic concepts of condensed
matter. We hope that our book will bring to them a flavor of live physics in
an informal manner as used by practicing researchers. It may help to open
their eyes to a more general fact that the science of physics often evolves
in its own ways quite different from those implied by the standard curric-
ulum textbooks, and, for some, to become an evidence that physics can be
relevant for PV.
Our next category of readers belongs to those in academic teaching and
research revolving around semiconductors, device physics, and thin film
structures. For that category, we aim to provide a fresh look at a number
xiv Preface

of subjects so established that they appear “carved in stone” for ages as


dictated by the classical photovoltaic science developed during the times
when thin film technology PV did not exist. Our book will invite that cat-
egory of readers to pay more attention to the issues of variability and sta-
tistics apparent to those working for industry rather than for academia.
Finally, we hope that some of our book topics and approaches can become
a part of university curricula for PV/device physics or related disciplines.
We identify our third category of readers as the industrial R&D profes-
sionals of all levels. Their business responsibilities and hectic environment
do not often leave much time to follow the published research and appreci-
ate new concepts. We believe that our book will help the curious readers of
that category to broaden their horizons and open new approaches towards
technology goals. On top of that, adopting some concepts of our book
could create beneficial synergy between physics and solar technology.

Victor Karpov
Diana Shvydka
Sylvania, Ohio, USA
February 2021
Part I
General and Thin Film PV

I. Introduction to Thin Film PV


A number of brilliant texts are readily available describing the general
photovoltaic (PV) principles and their classical implementations [1–6], as
well as various specifics of thin film PV [7–11], and we are not going to
duplicate them in any form. This section will briefly overview main book
concepts especially addressing the underlying physics of PV functionality
different between the crystalline and thin film devices. Such a synthetic
view will help to inter-relate various sides of the physics of thin film PV as
a subject of its own.

A. The Origin of PV. Junctions


We recall that PV effect in general is presented by the light generated voltage.
The PV voltage when the system absorbed light creates electric asymmetry
in the form of spatially separated opposite electric charges, electrons and
holes, which can be extracted and utilized in the form of electric current.
For the charge separation to occur, there must be a built-in electric field
as illustrated in Fig. 1 presenting a basic design of PV devices made of
two semiconductor layers. The built-in field originates from the junction
formed by those layers. Historically, such junctions were between p- and
n-types of semiconductor layers; hence, the name of p-n junctions tradi-
tionally associated with PV devices. However, p-n junctions are neither
necessary nor sufficient elements of systems with built-in fields as dis-
cussed below. For example, semiconductor/metal junctions in Fig. 1 will
generate their own built-in fields that can be either beneficial or detrimen-
tal to PV device functionality.
In general, the built-in electric fields always emerge with junctions of
any two chemically different materials. The underlying physics is that one
of the materials will be energetically more favorable for the electrons than
the other. To minimize the system energy, the electrons will therefore

Victor Karpov and Diana Shvydka. Physics of Thin-Film Photovoltaics, (1–12) © 2022 Scrivener
Publishing LLC

1
2 Physics of Thin-Film Photovoltaics

light
front contact

Layer 2

Layer 1
back contact

Fig. 1 Conceptual design of a solar cell. Front and back contacts are metallic, and the
former one is transparent to light (shown in waving lines). The presented built-in electric
field E is caused by the dark positive and negative charges shown as + and –. R is a load
resistor. For specificity, the diagram presents a two semiconductor layer design, such as
p- and n- materials with the field E in their junction proximity. However, sufficient electric
fields can exist as well in the proximities of semiconductor-metal junctions.

move there leaving the one with unbalanced positive charge; hence, the
built-in field between the spatially separated opposite charges of electrons
and holes, qualitatively similar to that of electric capacitor. The number of
electrons moving across the junction is determined by the balance between
the above mentioned energy gain and the energy loss due to the necessity
of overcoming the Coulomb attraction to the positive charges left behind.
The built-in electric fields of that nature are omnipresent and are not lim-
ited to photovoltaics, or p-n junctions, or other artificial structures.
Another reading of the latter statement is that photovoltaics do not nec-
essarily have to be related to or understood in terms of p-n junctions. In
reality, any (not only p-n) junctions of different materials produce built-in
electric fields. Some of them, but not all, create photovoltaic effect.
For example, a junction of two metals produces the built-in field under-
lying the phenomenon of thermoelectricity, but not suitable for PV because
the light does not penetrate in a metal deep enough and because that field is
screened (by metal electrons) beyond a nanometer thin region, insufficient
for light absorption. However the built-in fields of metal/semiconductor
(rather than metal/metal) junctions can make good diodes and PV devices.
As another example, we point at a charge acquired by a solid particle
immersed in a liquid. Curiously, that example explains how wines and
many other liquid products consist of charged micro particles suspended
in a somewhat ionized water. We will return to that example later in the
book describing the so called “red wine effect” in PV.
Note that junction fields require electronic exchange between two mate-
rials while their direct physical contact is not necessary. As an illustration,
General and Thin Film PV 3

chemically different metal electrodes of a capacitor will spontaneously


acquire opposite charges proportional to the difference in their work
functions. That process often called the ‘dielectric absorption’ takes time
sufficient for the inter-plate electron transfer. Along the same lines, two
different metals separated by a dielectric or semiconductor layer will
exchange electrons forming the built-in electric field throughout that layer
as described more in detail next (Sec. II).

B. Fundamental Material Requirements


A well known feature of all PV devices is that at least some of their consti-
tuting materials must have a not-too-wide forbidden gap, say G ~ 1–2 eV.
The origin of that criterion is that larger band gaps exceeding the energies
of most of sun spectra photons would not allow efficient light absorption
necessary to produce enough electrons and holes. However, there is also
the opposite requirement of those gaps being not too small, so that effi-
ciently absorbing narrow band semiconductors are not good for PV either.
As a result, semiconductors with substantial but not-too-wide forbidden
gaps G ~ 1–2 eV, such as Si, Ge, CdTe, CIGS and some others are most
suitable for PV.
The not-too-narrow gap limitation is less intuitive and may be worth
explaining here. It is dictated by PV functionality as a power source. The
power is a product of electric current and voltage, P = IV where both I and
V components must be not too small. I is proportional to the light induced
charge generation rate favoring significant absorption coefficients in the
sun spectral region of ħω ~ 0.5–2 eV. Forbidden gaps G ~ 1–2 eV of many
semiconductors fall in that region. To the contrary, the typical dielectric
gaps G ≳ 4 eV are too wide for sufficient absorption (for example, window
glass that is practically transparent and does not absorb light).
The requirement that undermines the suitability of narrow gap semi-
conductors (despite their strong absorption coefficients) is related to the
voltage component: V ≤ G/q, where q is the electron charge. To explain the
latter inequality we note that V can be related to the difference in Fermi
energies in two metal electrodes as illustrated in Fig. 2. We recall that the
Fermi (quasi-Fermi) energy describes the energy change due to removing
one particle from the system. Therefore, ∆EF ≡ EF1 – EF2 gives the energy
change due to transferring one electron between the two metals, and so
does qV. On the other hand, the metal Fermi energies must not overlap
with the semiconductor conduction or valence bands because such an over-
lap would mean that the semiconductor has in reality metal conductivity
(no gap between the Fermi level and forbidden gap edges). As illustrated
4 Physics of Thin-Film Photovoltaics

metal semiconductor metal

EF1

G eV

EF2

Fig. 2 Band diagram of a metal/semiconductor/metal structure where EF1 and EF2


represent quasi-Fermi levels determined by voltage drop V across the semiconductor.

in Fig. 2, G presents the upper limit of the difference between the elec-
tron and hole Fermi levels allowing their steady state spatial separation.
Increasing V beyond G/q would shift the Fermi levels beyond the forbid-
den gap turning the semiconductor into metal thus shorting the built in
field and the circuit.
Another wording of the same is that electrons and holes recombine very
efficiently when the two Fermi levels are close thus suppressing their spatial
separation and the built-in field. Assuming point defect mediated recom-
bination, it can be shown more quantitatively [12] that the electron-hole
recombination rate strongly accelerates with decrease of the gap G. As a
result, there exists a range of gaps optimizing PV performance, G ≲ 2 eV.
It should be remembered however, that the latter optimum gap prediction
was derived under certain assumptions about the nature of (defect facili-
tated) recombination, lack of traps, insignificant leakage due to shunts, and
others clearly outlined in the original work [12].

C. Charge Transport. Definition of Thin Film PV


We now consider charge carriers transport in the junction region where
the built in field is significant; that region is often called the space charge
region, simply because the field and charge density are E = 4πρ (where
E and ρ are correspondingly the field and charge density, and we use
the Gaussian system of units). The built-in field extends through a lim-
ited thickness lE determined by the condition that the change of electron
energy through lE equals the difference in Fermi levels ∆EF of the materials
involved. In the order of magnitude, the latter condition yields q 2nlE2 /ε ~ ∆EF
where n is the charge concentration (cm–3) and ε is the dielectric permit-
tivity (for simplicity, we neglect possible differences in n and ε between
the junction components). Using the ballpark values, ∆EF ~ 1 eV, ε ~ 10,
General and Thin Film PV 5
16
n ~ 10 cm –3 yields lE ~ 1 μm. This estimate is consistent with the pub-
lished data for various junctions.
In the built-in field region, the charge carrier transport is dominated
by drift with velocity v = μF where μ is the mobility and F = qE = qV/lE is
the electric force. The diffusion with coefficient D = μkT (according to the
Einstein formula where k is the Boltzmann constant and T is the tempera-
ture) is insignificant contributing to the transport time a relatively small
fraction of kT/qV ≪ 1 assuming the typical V ~ 1 V and kT ~ 0.025 eV
(room temperature).
The drift transport approximation is sufficient if the light absorption
is limited to the space charge region, which applies to many thin-film
structures. That approximation is well justified when the light absorption
length is short enough, labs ≲ lE, limiting charge carrier generation to the
built-in field region, which takes place in CdTe, CIGS, and a variety of
other thin-film PV materials. However, some crystalline structures, such
as Si and Ge, possess the non-direct band structures with absorption
lengths in the hundred of microns range. In such materials, charge carriers
generation extends way beyond the built-in region, and the transport is
dominated by diffusion rather than drift. We recall that the characteristic
diffusion distance and time are related as LD  Dt where D is the diffu-
sion coefficient.
The above discussed drift and diffusion times,

τE = l/v = l2/(μqV)    and    τD = l2/D, (1)

describe how long it take for a charge carrier to travel distance l to the
device terminals. These times must be short enough to avoid significant
electron-hole recombination. If the characteristic recombination time is
τr, we must require that the relevant transport time to device terminals,
τe or τd be shorter than τr. The corresponding criteria are discussed in
Sec. XVI below. Here, it suffices to note that τr is extremely sensitive to
material parameters and device structure, which brings in multiple ele-
ments of materials science and PV technology. In particular, recombina-
tion processes are strongly facilitated by structural defects (imperfections,
impurities, surface states) that depend on details of material composition;
interfacial recombination can play a significant role for some junctions, etc.
At a qualitative level, we discriminate between thick and thin PV
structures. Typical of crystalline PV, the former are such that the elec-
trons and holes diffuse long distances to the device terminals and τd can
be longer than τr, i.e. interaction with defects becomes inevitable during
their long trips. To the contrary, the characteristic drift time τe is much
6 Physics of Thin-Film Photovoltaics

shorter than τr in thin PV where recombination becomes almost immate-


rial. As a result, defects and imperfections determine thick PV efficiency
quite significantly. That observation undisputable for thick PV, can become
confusing when uncritically extended over thin film PV where it is not
warranted. Furthermore, a century-long successful history of crystalline
(Si, in the first place) PV made the concept of defect facilitated recombi-
nation literary a commandment dominating all PV understanding across
the discipline. That paradigm has triggered various defect characterization
studies in thin film structures, and, while no correlations between defects
and thin film PV performances was established, that activity continues.
Of course, the difference between ‘thick’ and ‘thin’ devices is related to
the fact that the former have thickness l ≫ lE where lE is the linear dimen-
sion of the built-in field region, as presented in Fig. 3. To the contrary, l ≲ lE
or l is just slightly greater than lE for ‘thin’ devices. In the absence of field,
charge carriers have to travel by diffusion. Taking into account the Einstein
relation between the mobility and diffusion coefficient, the ratio of ‘thick’
over ‘thin’ device transport times is estimated as

thick Vq l 2
 1 (2)
thin kT lE2

where V is the voltage across the device, q is the electron charge. Given
the typical |Ve| ~ 1 eV, kT ~ 0.025 eV (room temperature), l on the order of
hundreds of micron and lE in microns, the ratio in Eq. (2) can be as large as
5 6
10 –10 . The latter inequality is consistent with the typical experimental

lE l

Fig. 3 Diffusion of the photogenerated electrons and holes to the electrodes in a ‘thick’
device. White circles represent defect recombination centers. The figure is aimed to show
that interactions with such centers may be likely when charge carriers exercise long-time
diffusion movements.
General and Thin Film PV 7

data: the diffusion times τd,thick ~ 10–100 μs (in crystalline Si PV) and td
~ 1–10 ns in thin film PV. With those numerical values, using the typical
recombination times τr in the range of micro- to milliseconds (as evaluated
for various semiconductors [3, 13–17]), one should expect τr ≫ τd for thin
film PV.
Based on the latter estimate, thin film PV can operate in the opposite
limit τ ≪ τr where the role of defect related recombination is on average
insignificant. Unfortunately, they have other vulnerabilities related to lat-
eral nonuniformity of their parameters. From a very general point of view,
such nonuniformity originates from the lack of transversal self-averaging
across too thin structures, which allows significant fluctuations between
different spots. Due to device thinness, some sparse spots represent para-
sitic conductive paths (ohmic or non-ohmic) between opposite electrodes.
These paths work as recombination highways degrading PV efficiency. We
emphasize that such a nonuniform recombination is qualitatively differ-
ent from that by point defects in thick PV and call for different mitigating
strategies.
Furthermore, having reached the electrodes their respected ter-
minals, electrons and holes undergo fast lateral transport inside the
contact metals taking them to the above mentioned recombination
channels, through which they disappear. We will see in what follows
that the recombination times τrc of such channels (presented in Fig. 4)
form an exponentially broad spectrum decaying towards shorter val-
ues. The channels with characteristic recombination times of the order
of charge carrier travel times τ dominate because the channels with τrc
≫ τ are inefficient, while those with τrc ≪ τ are too rare. This feature is
more quantitatively explained in Sec. XVI following Fig. 104 and subse-
quent chapters. At this point, we conclude that thin film PV can operate
through multidimensional processes where electric currents across and
along films are strongly coupled. From the practical perspective, surface
treatments that can block the recombination channels appear to be a
strategy improving thin film efficiency. That approach has been success-
ful as established on purely empirical grounds before it was theoretically
substantiated.

D. Distinctive Features of Thin Film PV


Another distinctive feature of thin film PV is that their forming layers are
noncrystalline, but rather polycrystalline (CdTe and CIGS types of PV)
or amorphous (a-Si:H based PV). That type of ’imperfect’ morphology is
8 Physics of Thin-Film Photovoltaics

lE ~ l

Fig. 4 In thin film devices, drift rather than diffusion can determine the kinetics of
photogenerated carriers. They readily arrive at the corresponding electrodes avoiding
interactions with recombination centers. Having reached the electrodes they can move
to the entrances of rare but highly efficient recombination channels shorting between the
electrodes and playing the role of recombination highways.

inevitable with inexpensive fast deposition techniques not giving enough


time for satisfactory crystallization. The polycrystalline films consist of
tightly packed individual grains, each having more or less perfect crys-
talline order inside. They are formed during deposition and subsequent
treatments where random atomic configurations are ordered within local
regions dictating grain sizes. In the course of such crystallization, material
pushes away all ’foreign’ atoms and structural units towards grain bound-
aries (GB). (Similar to apples in a freezer where water crystallization breaks
and pushes away organic tissues.) As a result, GB material is chemically
different from that of interior. These chemical differences entail built-in
electric fields that will spatially separate electrons and holes between the
grain interior and GB (see Fig. 5) creating local electric field and poten-
tial variations. One other general consequence of independently forming
grains is that having grown enough they compete for space and exert pres-
sure on each other upon physical contact. The resulted compressive stress
can translate into electric potential when the material structure is piezo-­
active, such as with CdS, ZnO and some others, as explained more in detail
next (Sec. XVIII).
As illustrated in Fig. 5, spatial separation of charge carriers at GB can
significantly decrease the probability of their recombination compared
to crystalline materials where electrons and holes are distributed uni-
formly. That creates a possibility for polycrystalline thin film PV to poten-
tially outperform their crystalline counterparts. On the other hand, high
concentration of impurities and defects in the grain boundary interfaces
enhances the recombination processes, which can make grain boundar-
ies detrimental and polycrystalline material PV devices perform worse
General and Thin Film PV 9

e
EC

G
EF

EV
VB
h

Fig. 5 Role of grain boundaries in noncrystalline thin film PV. Left: Sketch of the
electron hole pair creation by a photon of energy exceeding the forbidden gap G. Ec and
Ev are conduction and valence band edges, EF is the Fermi energy. Short horizontal lines
represent defect states located at the grain boundary. Note that the positive (attractive
to electrons) GB charge is assumed arbitrarily. For the case of negative GB charge, the
potential well (barrier) for electrons (holes) will turn into a barrier (well). Right: the GB
barrier requires charge carrier activation [with probability proportional to exp(–VB/kT)]
which suppresses recombination (presented by vertical arrows).

than crystalline PV. The latter feature was evidenced in some polycrystal-
line Si structures. Furthermore, it will be shown below (Sec. XVI) that the
very nature of recombination processes in non-crystalline materials can
be significantly different from that dominating crystalline PV and affect-
ing both the GB and main junction processes. Overall, the role of grain
boundaries remains debatable and material/recipe specific with experi-
ments exhibiting either beneficial or detrimental GB effects. It is possible
that some of GBs in thin films form, with certain probability, much more
effective recombination channels than others. Those channels can play the
role of recombination highways depicted in Fig. 4.
A physically nontrivial and practically important property of thin film
PV (further discussed in Sec. V) is that, under certain conditions, they
exhibit thermal runaway instabilities favored by low heat transfer param-
eters as well as low sheet resistances along with significant currents and
voltages. These instabilities result in local (~ 1 mm) spots of elevated tem-
perature that can degrade PV performance and trigger nonuniform struc-
ture damage. We will discuss later their underlying physics and factors that
can be tweaked to mitigate the latter detrimental effects by proper device
engineering.
A more specific structure of the two major brands, CdTe and CIGS
based thin film PV is presented in Fig. 6. It is understood that in reality
the glass is facing up in CdTe cells. The CdTe and CIGS layers are absorb-
ers responsible for the light conversion into photogenerated electrons and
holes. The question of their doping has been debatable for decades, with
10 Physics of Thin-Film Photovoltaics

Grid
Glass
Buffer (ZnO)
TCO
CdS
CdS
CuIn(Ga)Se2
CdTe
Mo

Back metal Glass

(a) (b)

Fig. 6 A cross-sectional view of a superstrate (glass sheet up) CdTe based and substrate
(glass sheet down) CIGS based thin film solar cell (not to scale). The CdTe cell front
contact is formed by the transparent conductive oxide (TCO). The typical layer
thicknesses are 2-4 µm for CdTe and CIGS, 0.1-0.2 µm for CdS, 0.5 µm for TCO and
buffer layer, 0.1-0.2 µm for back metal (Mo for CIGS), 1-3 mm for glass. The drawing
does not show the polycrystalline structure of CdTe, CIGS, and CdS layers. Red arrows
represent incident sunlight.

extreme views ranging from that these materials cannot be effectively


doped to that of beneficial or detrimental role of certain impurities. Many
of such mutually exclusive points have found experimental confirmations
with particular device recipes.
Furthermore, the role of CdS layer often found with CdTe and CIGS
remained mysterious long after it was empirically found to be extremely
important and practically irreplaceable in spite of its lack of contribution
to photocurrent. We shall see in what follows (Sec. XVIII) that its piezo-
electric properties may be a key leading to piezo-photovoltaic coupling
beneficial for PV operations.
Finally, the back contact layer renders yet another set of puzzles. While
the Mo back contact is typical of CIGS PV, there are several successful rec-
ipes of back contact in CdTe based PV. It was found indeed that the back
contact has profoundly strong effect on device operations in spite of the
fact that it is not photoactive (light does not penetrate there). The so-called
back barrier due to the junction field between the semiconductor and back
metal turns out to be a culprit. More in detail, the physics of back contact
functionality will be discussed in Sec. VB. The front contact layer, such as
TCO in CdTe based PV or ZnO in CIGS, has its nontrivial properties as
well, as also discussed in the subsequent chapters.
Last but not the least is the issue of thin film PV degradation. Light,
elevated temperatures, and humidity can all play a detrimental role in
General and Thin Film PV 11

degradation processes. Some PV companies went under being unable to


solve these problems. It often happened that the most efficient PV devices
exhibited the highest degradation rates calling upon tradeoffs between
efficiency and stability. It should be understood that PV degradation is
inevitable with any PV in full compliance with the general Le Chatelier -
Brown principle stating that when a settled system is disturbed, it will
adjust to diminish the change that has been made to it, or, roughly stated,
any change in status quo prompts an opposing reaction in the responding
system [18]. With respect to PV, that principle predicts that PV structures
will evolve in such a way as to decrease the light induced electric currents;
hence, degradation. While all PVs degrade, the thin film ones are espe-
cially vulnerable due to their thinness allowing formation of conductive
transverse channels (shunts). In addition, thin film structures accumulate
extremely high electrostatic energy density (similar to narrow gap (~ 1
µm) electric capacitors) capable of discharging with huge power dissipa-
tion transforming material structure towards electric shorts that minimize
the electric energy. A more in detail discussion of these issues will be given
in Sec. XXVII B.

References
1. M. A. Green, Solar Cells: Operating Principles, Technology and System
Applications, Prentice Hall, Englewood Cliffs, N.J., 1982.
2. A. L. Fahrenbruch, R. H. Bube, Fundamentals of Solar Cells, Academic Press
(1983).
3. J. L. Gray, The physics of Solar Cells, in Handbook of Photovoltaic Science and
Engineering, p. 61, Edited by A. Lique and S. Heggedus, Wiley (2002).
4. J. Nelson, The Physics of Solar Cells, Imperial College Press (2003).
5. P. Wurfel, Physics of Solar Cells, Wiley (2005).
6. T. Dittrich, Material Aspects of Solar Cells, Imperial College Press (2015).
7. T. J. Coutts, J. S. Ward, D. L. Young, K. A. Emery, T. A. Gessert, and R. Noufi,
Critical Issues in the Design of Polycrystalline, Thin-film Tandem Solar
Cells, Prog. Photovolt: Res. Appl. 11, 359 (2003).
8. Thin Film Solar Cells Fabrication, Characterization, and Applications, Edited
by J. Poortmans and V. Arkhipov, Wiley (2007).
9. S. R. Kodigala, Cu(In1xGaxSe2 Based Thin Film Solar Cells Elsevier,
Amsterdam New York, (2010).
10. R. Scheer and H.-W. Schock, Chalcogenide Photovoltaics, Wiley (2011).
11. Advanced Characterization Techniques for Thin Film Solar Cells, Edited by
D. Abou-Ras, T. Kirchartz, and U. Rau, Wiley (2011).
12 Physics of Thin-Film Photovoltaics

12. W. Shockley and H. J. Queisser, Detailed Balance Limit of Efficiency of p-n


Junction Solar Cells, J. Appl. Phys., 32, 510 (1961).
13. A. G. Milns, Deep impurities in semiconductors, Wiley Interscience (1973).
14. D. K. Schroder, Carrier Lifetimes in Silicon, IEEE Trans. Electron Devices,
44, 160 (1997).
15. S. Parola, et al, Study of photoluminescence decay by time-correlated single
photon counting for the determination of the minority-carrier lifetime in sil-
icon, 4th International Conference on Silicon Photovoltaics, SiliconPV 2014,
Energy Procedia 55, 121 (2014).
16. S. Khatavkar et al, Measurement of Relaxation Time of Excess Carriers in
Siand CIGS Solar Cells by Modulated Electroluminescence, Technique,
Phys. Stat. Solidi. A, 215, 1700267 (2018).
17. V. N. Abakumov, V. I. Perel, I. N. Yassievich, Nonradiative Recombination in
Semiconductors (Modern Problems in Condensed Matter Sciences), North-
Holland (1991).
18. Wikipedia paper https://en.wikipedia.org/wiki/Le_Chatelier%27s_principle
Part II
One-Dimensional (1D) Diodes and PV

II. 1D Diode
The concept of diode underlies almost all the aspects of PV operations.
Diode means a two-terminal device allowing an electric current to pass
in one (forward) direction, while blocking it in the opposite (reverse)
­direction. Its nonlinear current-voltage (IV) characteristic is omnipresent
in structures where the electron transport is affected by potential barriers.
In reality, those barriers are associated with junctions. Here we discuss the
simplest case of such a barrier to see how it leads to the diode type IV and
which barrier parameters are important. In this part of the book, we con-
sider only 1D structures and processes implying that there is no significant
lateral nonuniformities and currents.
It has become a standard in diode introduction literature to start with the
notion of p-n junction [1–11]. However important, the latter concept appears
somewhat oversold in applications to thin film PV at least. We would like to
emphasize that the rectifying diode characteristics can be caused by practi-
cally any kind of junctions. Therefore, we decided to start with an example
far away from the p-n situation to show how the diode IV becomes possible
otherwise. We will touch upon the p-n junction concept later referring in the
meantime to the ample high-quality literature presenting that concept.

A. Metal-Insulator-Metal Diode
We consider a barrier formed by a contact of chemically different mate-
rials, so that the electron energy changes at their interface. The material
where it is higher represents a potential barrier that the electron needs to
overcome in order to penetrate that material. A barrier between a metal
and vacuum represents the simplest case. By definition, its height equals
the metal work function, i.e. the minimum energy required to extract an
electron from the metal to the vacuum. In other words, it is the difference
between vacuum level of energy and the Fermi level in the metal.

Victor Karpov and Diana Shvydka. Physics of Thin-Film Photovoltaics, (13–78) © 2022 Scrivener
Publishing LLC

13
14 Physics of Thin-Film Photovoltaics

Fig. 7 represents the energy band diagram of a nonmetal layer sand-


wiched between two identical metal electrodes, the simplest junction sys-
tem allowing connections to the external power source. Since the Fermi
energies in the latter lie against the forbidden gap of the middle layer, the
lowest available energy states are above that layer conduction band edge.
The electrons in metals have to increase their energies, from the Fermi lev-
els up to that edge, in order to penetrate into the nonmetal layer. In other
words, they have to overcome the corresponding potential barriers W1 or
W2 in Fig. 7. Note that W1 = W2 for a particular case of similar metals in
Fig. 7. Note also that neither W1 nor W2 are the metal work functions. They
represent instead the differences between the metal work functions and
electron affinity of the non-metal layer. The affinity is defined as a mini-
mum work necessary to extract the electron from the edge of the conduc-
tion band to the vacuum level (which is assumed to lie above that edge).
Shown in Fig. 8, the current-voltage characteristic for a symmetric struc-
ture in Fig. 7 demonstrates saturation for strong enough voltages of both
signs. The underlying physics is that the current approaches a maximum
when the two barriers differ enough.
A comment is in order with regards to the diagrams in Figs. 7, 9 and
many others throughout the book: they all represent coordinate depen-
dencies for the electron energy. In some cases, it is customary to use the
terminology of the electric potential dividing the energy by the elemen-
tal charge q. It must be remembered however that the latter terminology

J1 J2
J1 J2

W1 W2 W1 W2+qV
EF1
EF1 EF2 qV
EF2

(a) (b)

Fig. 7 Energy band diagram of a symmetric potential barrier separating two metal
electrodes. The barrier heights W1, W2 measured from the corresponding Fermi levels of
the metals EF1, EF2 are the same when no voltage is applied, and the currents from these
metals balance each other. Applying voltage V shifts the Fermi levels by the energy qV
relative to each other, causing the corresponding difference in the current.
One-Dimensional (1D) Diodes and PV 15

J, arb. units
V, arb. units

kT/q

Fig. 8 Current-voltage characteristic of a system with symmetric barrier illustrated in Fig. 7.


The saturation current at both negative and positive biases is equal to J00 exp(−qW1/kT).

J1 J2
EF1

W2 EF1
qV
EF2 EF1 EF2 qV EF2 EF2
V=0
EF1 W1 -qV

(a) (b) (c) (d)

Fig. 9 Energy band diagram of an asymmetric potential barrier between two different
metal electrodes. (a) The current J2 ≫ J11 is the saturated current under negative (reverse)
bias. (b) The currents are equal, J1 = J2 under zero bias as limited by activation over the
barrier W2. (c) The forward current J1 exponentially grows with further increase in bias
and approaches saturation close to the flat band barrier condition where J2 ≪ J1.
(d) Further increase in forward bias exponentially suppresses the reverse current J2.

(often used in physics books) is not exactly consistent with the definition
for the electric potential as the energy per charge: it neglects the negativity
of the electron charge producing strictly speaking minus of the electric
potential instead of the potential itself. For example the energy of a positive
particle will increase in the direction where the electron energy decreases.
In most cases the above mentioned negligence does not result in any mis-
understanding. We encourage the reader to keep this comment in mind
while switching between the physics related diagrams and electric poten-
tial mappings such as presented in Figs. 116 and 117. We assume that the
barrier is surmounted by activation i.e. the necessary energy is provided by
16 Physics of Thin-Film Photovoltaics

thermal fluctuations. The probability of such fluctuations is proportional


to the Boltzmann exponential exp(−W1(2)/kT). It is typical that W1(2) >> kT.
For example, kT ≈ 0.026 eV at room temperature while W1(2) is on the
order of characteristic differences in band energies in solids, i.e. W1(2) ≲ 1
eV. With the above factor in mind, the current density (per area) from a
metal to nonmetal layer can be presented as

W1(2)
J1(2) J 00 exp .
kT

The preexponential factor J00 will coincide with the well-known preex-
ponential of the thermoelectronic emission [12],
4 πqmk 2 A
J 00 = A∗T 2 , with A∗ = 3 ≈ 120 2 2 (3)
h cm K
where A* is the Richardson constant. Its numerical value here is given for
the case of free electrons when the effective mass m equals the true electron
mass; h is the Planck’s constant. Eq. (3) is derived by counting the number
of electrons crossing unit area per time, nv 4 with v ⊥ = 8kT πm being
the average velocity component perpendicular to the area element, and
3/2
n the electron concentration. An additional power of temperature, T ,
appears due to the concentration of states available to the electrons in a
narrow energy band kT. That concentration is roughly estimated as 1/λ3
where h mkT is the de Broglie wavelength of the electrons.
In many cases, Eq. (3) can be modified to account for diffusion and the
details of band structure. It is important however, that in all such cases J00
remains a power function of temperature that is much weaker than the
exponential multiplier in the equation for current. In a good approxima-
tion, the preexponential can be treated as a constant and the numerical
value in Eq. (3) gives a ballpark of expected preexponential multipliers.
A word of caution may be in order here to stipulate that, for simplic-
ity, the non-metal layer is tacitly assumed to make negligibly small effect
on the electric current. In other words, its resistance is taken to be much
below that of the junction itself: it is much ‘easier’ for the electron to drift
through that layer than to get activated into it through the junction.
Consider next the ‘dark’ (i.e. no illumination) current-voltage charac-
teristics (neglecting photo-generated carriers). The energy difference W1(2)
between the barrier height at the interface and the Fermi level EF1(2) in the
adjacent metal remains constant under external field as fixed by the differ-
ence in parameters of the two materials.
One-Dimensional (1D) Diodes and PV 17

When the electrodes are identical as shown in Fig. 7, electric bias V cre-
ates the difference qV between their Fermi levels. [We recall that the Fermi
level coincides with the chemical potential equal to the change in system
energy when the number of electrons changes by one. On the other hand
moving the electron across the electric potential difference results in the
energy change qV.]
While the energies W1 and W2 remain constant, the activation energy
for the electrons of one of the electrodes increases by qV in Fig. 7, since
they have to overcome the maximum barrier located at the opposite inter-
face. The difference qV in the left and right barriers makes the correspond-
ing currents different. For the case of positive voltages illustrated in Fig. 7,
their absolute values can be written as

J1 = J00 exp(−qW1/kT) and J2 = J1 exp (–qV/kT)

and the current voltage characteristics, J = J1 – J2 takes the form

qV
J J1 1 exp . (4)
kT

It saturates as illustrated in Fig. 8.


While the IV characteristic in Eq. (4) is non-ohmic, it does not yet rep-
resent a major diode property of exponentially strong asymmetry between
the currents of opposite senses. That property naturally appears with
asymmetric barrier illustrated in Fig. 9 where the two interfacial energy
steps W1 and W2 are substantially different. Physically, this system can be
a dielectric or semiconductor layer sandwiched between two different met-
als, for example, CdTe between Al and Au. In the asymmetric case, the
barrier shape exhibits a finite slope even under zero bias as shown in Fig.
9(b). According to Fig. 9, the barrier heights limiting the current are given
by the effective values

W1,eff = max{W1, W2 − qV}, (5)


W2,eff = max{W2, W1 + qV}.

The corresponding currents are given by,

W1(2),eff
J1(2) J 0 exp .
kT
18 Physics of Thin-Film Photovoltaics

They have exponentially different saturated values:

J max W2 W1
exp .
J min kT

In Fig. 10 these currents correspond to the regions (d) and (a) ­respectively.
The voltage region (c) of increasing forward current is by the factor
(W2 − W1)/kT ≫ 1 wider than the region (b) of increasing reverse current
corresponding to the diagrams (c) and (b) in Fig. 9.
The IV curve of Fig. 10 represents the typical diode characteristic. We
note that the saturation-like region under forward bias region is always
present in real diodes, although it is often omitted in semiconductor
textbooks as irrelevant. Furthermore that region per se is tantamount to
non-dissipative current neglecting the inevitable ohmic resistance in a
material. Taking the latter into account renders that region a finite slope as
depicted in Fig. 10 by dashed line.
Limiting ourselves to the regions (a) -(c), the current voltage character-
istic can be now presented as

qV W2
J J 0 exp 1 , J0 J 00 exp . (6)
kT kT

Ohmic
resistance
Joexp(-W1/kT) (d)
J, arb. units

(c)

(b) (W2-W2)/q
V, arb. units (a)

kT/q Joexp(-W2/kT)

Fig. 10 Current voltage characteristic of a system with asymmetric barrier illustrated in


Fig. 9. Regions (a) - (d) refer to the corresponding band diagrams in Fig. 9. The saturation
current in the reverse and forward bias regions are J1 and J2 ≫ J1 respectively.
One-Dimensional (1D) Diodes and PV 19

The latter is the standard diode equation containing two important fac-
tors: strong asymmetry between reverse and forward regions and expo-
nentially low reverse saturation current. We would like to emphasize that
asymmetric potential barrier – not necessarily of rectangular or trapezoid
shape – is a single major factor leading to the IV characteristics of diode
type. Because such barriers are omnipresent (their absence would require
rather special circumstances), the diode IV are found everywhere in elec-
tronics, device physics, and even in biophysics where they exhibit them-
selves e.g. as current voltage characteristics of ion channels in biological
membranes [13].
We note finally that our analysis here tacitly neglected a number of pos-
sible complications, such as interfacial and metal induced gap states, possi-
ble screening, localized states and recombination in the intermediate layer,
which will affect the barrier heights and the saturation current J0, without
any impact on the functional form of the current voltage characteristic in
Eq. (6).

B. Schottky, Reach-Through, and Field-Compensation


Diodes

1. Schottky Diode
Schottky diode is in general due to a contact between a metal and a semicon-
ductor. Metal contacts are inevitable with all semiconductor applications.
Because the Fermi levels of metals and semiconductor components are dif-
ferent, the electrons move from one of them to another forming excessive
electric charges across the junction as illustrated in Fig. 11. For the ideal
case of perfectly clean uniform interface, the classic Mott-Schottky theory
predicts the interfacial barrier height W equal the absolute value of the
Fermi levels differential. The characteristic width of that barrier is given by
the electrostatic screening length, which is often described for electrons Lsn
and holes Ls(p) by the equation (in Gaussian units)


Lsn( p ) = (7)
2πN d (a )e 2

where ε is the dielectric permittivity, and Nd(a) stands for the concentration
of donors (acceptors) in respectively n-type or p-type materials. As a ball-
15 −3
park estimate, one can assume W ~ 1 eV and Nd(a) ~ 10 cm , and ε ~ 10,
which yields Lsn(P) ~ 1 μm.
20 Physics of Thin-Film Photovoltaics

Eq. (7) assumes discrete energy levels of donors or acceptors responsible


for the electric charge in the band bending region of Fig. 11. To the contrary,
when the material is non-crystalline and has a quasi-continuous spectrum
of energy levels with density g(EF) (cm−3erg−1) at the Fermi energy in the
forbidden gap, the screening length is described by the equation [14]

ε
Ls = 2 . (8)
4 πe g ( EF )

15
Using for a rough estimate the same ~ 10 cm−3 localized states more
or less uniformly distributed over the energies in the gap of ~ 1 eV yields
15 −1
g(EF) ~ 10 cm−3eV , and again Ls ~ 1 μm. The characteristic Schottky
barrier width in the micron range appears to be a rule of thumb estimate
for various materials.
While the Mott-Schottky theory remains a classical concept presented
in every semi-conductor textbook, more recent studies revealed a rather
complicated picture [15] with no general rules for the Schottky barrier

(a) (b)

Fermi level
energy

Fermi level Fermi level

Metal Semiconductor Metal Semiconductor

(c) (d)
Conduction band

Barrier
Semiconductor for holes
energy

Valence band

Fig. 11 Schottky barrier formation: (a) No electric contact; the Fermi level is arbitrarily
chosen to be higher in a metal. (b) When the Fermi energies of two materials put in
contact level out, the metal and semiconductor acquire opposite excessive charges. (c) The
positively charged metal repels holes and attracts electrons; (d) Tantamount to a barrier
for holes.
One-Dimensional (1D) Diodes and PV 21

conduction band

forbidden gap
∆W
EF
ψ(x)

∆x
x
metal insulator

Fig. 12 Gap states due to electron tunneling from the metal to dielectric side create
electric dipoles with charges presented by ⊕ and ⊖ symbols. ψ(x) stands for the electron
wave function of the electron state in a metal facing forbidden gap in the tangent
dielectric.

height. The existing models accounting for surface and metal induced gap
states [16] along with experimental results [17, 18] are not consistent with
the Mott-Schottky concept of simple electron transfer (built-in field for-
mation) between the two materials at MS interfaces.
It is quite obvious that interfacial localized states on the metallurgical
junction can hold electric charges creating additional contribution to the
barrier hight and width. However, even with the ideal interfaces, the gap
state phenomenon illustrated in Fig. 12 predicts substantial deviations
from the Mott-Schottky theory. Its underlying physics is that the electron
states below the Fermi level EF on a metal side penetrate by tunneling
into the forbidden gap of an adjacent semiconductor under the barrier of
height ∆W. The characteristic penetration length Dx   mDW is on
the order of several angstrom for the typical parameters ∆W ~ 1 eV, and
−27
electron mass m ~ 10 g. Along with the left behind positive charges, these
electron tails create an electric dipole layer. Its corresponding energy dif-
ferential is estimated as q 2 εDx  DWq 4 (m  2 ε 2 )  0.1 − 1 eV . Because
the latter value is comparable to the original Schottky-Mott barrier of
|W1 — W2| we conclude that other factors (forbidden gap, electron effec-
tive mass, surface states) may have very significant effect on the barrier
height making it more an empirical parameter than a predicted value.

2. Reach Through Diodes


When the semiconductor thickness is large enough compared to the
screening length [of Eq. (7) or (8)], the two metal contacts on its opposite
surfaces can be considered independent of each other. Such a structure is
sketched in Fig. 13 where the two metals are different causing asymmetric
22 Physics of Thin-Film Photovoltaics

We2
We1 We2 We1 EF2
G
qV Wh2
EF1

Wh1 Wh2 Wh1


Ls2 Ls2
Ls1 Ls1
(a) (b)

Fig. 13 Reach-through structure of a semiconductor layer sandwiched between two


different metals when the layer thickness exceeds the screening length.

band bending. The parameters assumed (arbitrarily here) are such that the
energy barriers involved form the hierarchy Wh2 < Wh1 < We1 = We2 at zero
voltage in Fig. 13 (a). For small voltages V applied as shown in Fig. 13 (b),
the current is due to hole transport remains constant (electronic barriers
are too high and block the electrons). The current is then determined by
the barrier Wh2. However, from some voltage on, the electronic current
controlled by the barrier We2 starts contributing exponentially strong, i.e.
the total current sharply increases, which phenomenon is often referred to
as the ‘reach-through’. Various other band diagrams can lead to the reach-
through behavior [12]. The corresponding current voltage characteristics
are presented in Sec. VB below with PV applications.
The voltage Vrt upon which the double Schottky barrier structure turns
into the reach-through regime can be estimated from the condition that
the maximum energy point in the diagram of Fig. 13 (a) shifts all the way
through the system forming the configuration of Fig. 13 (b). Based on the
(possibly oversimplified) model of uniformly doped material [19],

2Wh1 Wh2
VRT 1 . (9)
q Wh1

Furthermore, using the available data and assuming that the electric
current through the barrier Wh2 is thermally activated, it was estimated
that Wh2 ~ 0.5 V. Because Wh1 ~ 1 eV, we finally get a rough estimate Vrt
~ 0.6 V.
The most sensitive approximation behind Eq. (9) is that it completely
neglects electron tunneling. The latter is more difficult to evaluate for it
depends on the barrier shape, generally unknown. Also, such tunneling
can be thermally and defect assisted, which makes it laterally nonuniform.
It will be discussed more in subsequent chapters.
One-Dimensional (1D) Diodes and PV 23

The reach-through related issues will be further discussed in connection


with the back contact problems and related nonuniformities (Sec VB) and
Schottky barrier nonuniformities (Sec IX).

3. Field Compensation Diode


Field-compensation diodes possess the property that the semiconductor
charge carrier concentration decreases due to the metal contact related
built-in electric field. The underlying physics is somewhat similar to that of
self-compensation phenomenon (see e.g. Ref. [20] and references therein)
where a semiconductor spontaneously generates deep level defects that
trap the charge carriers thus decreasing the system energy when that
gain exceeds the energy cost of defect creation. A unique feature of the
field-compensation is that the energy decrease is achieved through its elec-
2
trostatic component ∫ dx(εE /8π) by suppressing the built-in field via com-
pensation as illustrated in Fig. 14. That effect originally analyzed for a more
complex PV structure [21], has similarity with that of the field-induced
doping [22].
We now use a simple approximation to establish the condition for that
effect. We consider both screening lengths in Fig. 14 the same and equal
Ls. We approximate the built-in fields as E1(2) = W1(2)/qLs and their related
2
energy as WE = ε( E1(2) / 8π )ALs where A is the contact area. Let wd denote
the energy spent to create a deep level defect responsible for compensation.
(Assuming, in the spirit of self-compensation mechanism, some energy
gain due to the charge carrier localization, wd would stand for the differ-
ence between the former energy loss and the latter gain. In case the so
defined Wd is negative, the compensation takes place regardless of the field
effect.) The energy loss for defect generation is estimated as Wd = NLAwd
where L is the thickness of the semiconductor layer and N is the defect con-
centration. Neglecting the final field energy in the compensated structure

We1 G
We2
L
L G

Wh2
Wh1 Ls2
Ls1
(a) (b)

Fig. 14 Metal-semiconductor-metal structure before field compensation (a) and after that
(b). Note the decrease in Fermi energy and increase in screening length making the field
almost uniform.
24 Physics of Thin-Film Photovoltaics

compared to WE and omitting all the numerical multipliers of the order of


unity, the field compensation criterion WE/Wd > 1 becomes,

W12(2)ε
> 1. (10)
8π wd q 2 NLL s

15 −3
For a rough numerical estimate we use W1(2) ~ wd ~ 1 eV, N ~ 10 cm ,
L ~ Ls ~ 3 μm, ε ~ 10, which yields the left-hand-side of Eq. (10) on the
order of one. We conclude that the field compensation effect is possible
provided favorable parameters in thin film structures. It seems to be unre-
alistic for semiconductor layers thicker than 10 μm and would favor flex-
ible structures, such as chalcogenide based CdS, CdTe, CIGS, where the
characteristic energies wd of defect creation can be relative low.
The field-compensation effect can result in the extension of the built-in
electric field through out the structure, which improves the carrier
­collection. More important can be its caused communication between
the two layers tangent to the opposite semiconductor surfaces in Fig.
14 through the suppression of screening. That opens a avenue to govern
device operations by attaching a desired complementary layer across one
of the semiconductor components. Such can be e.g. a buffer layer on the
opposite side of CdS component of CdS/CdTe PV structures. There are
numerous experimental evidences in favor of such interpretation [21].

C. P-N Homo-Junctions
Fig. 15 presents the standard diagrams for p-n homojunctions commonly
adopted to describe the properties of crystalline PV. We note that in the
doped crystalline material the electrostatic screening lengths are pre-
sented by Eq. (7). The corresponding diffusion lengths LDn Dn n and
LDp D p p with Dn(p) and τn(p) being the electrons (holes) diffusion coef-
ficients and lifetimes, determine how far the charge carriers can diffuse
during their lifetimes. As usual, the quasi-Fermi levels EFn( p) in Fig. 15
are defined in such a way that the nonequilibrium charge carrier concen-
trations n(p) are proportional to exp(EFn( p) kT ) mimicking the relation
n0(p0) ∝ exp(EFn(p)/kT) between the Fermi energy EFn(p) and equilibrium
charge carrier concentration n0(p0).
One feature misrepresented in Fig. 15 is a possible strong asymme-
try between the n- and p-parts of a diode. For illustrative purposes, we
20 −3 15 −3
mention here3, Nd ~ 10 cm vs. Na ~ 10 cm , Lsn ~ 0.35 μm vs. Lsp ~
One-Dimensional (1D) Diodes and PV 25

Real space
p-type n-type

Lsn

LDp Lsp LDn

Energy space

quazi-Fermi level
Fermi level

Fermi level
quazi-Fermi level

Fig. 15 The real space and energy space diagrams for the classical p-n homojunction
under light. Lsn and Lsp are the electrostatic screening lengths in n and p regions, so
that Ls = Lsn + Lsp is the depletion width. The red arrows represent light and its e – h
generation process. The depicted relations between device linear sizes, screening lengths,
and diffusion lengths are not to scale and rather arbitrary in the drawing.

300 μm, and LDn ~ 12 μm vs. LDp ~ 1100 μm corresponding to τn ~ 350


μs and τn ~ 1 μs.
In connection with the latter numerical parameters, we would like
to note that the highly doped n-type material becomes rather imper-
fect containing significant concentrations of various defects facilitating
­recombination. It is natural then that the lifetime in n-material is so much
shorter than that in p-type. Also, we observe that the lifetimes in good
(p-type) and defective (n-type) materials vary in the range of roughly 1
μs to 1 ms, a ballpark for Si and many other semiconductors [23–27, 30].
An important feature of the classical p-n model (not presented in Fig.
15) is that it can include surface recombination implying that because of
the relatively high degree of structural disorder in the interfaces, they can
serve as efficient recombination facilitators. The surface facilitated recom-
bination can be described by the corresponding lifetime τs that contrib-
utes to the charge carrier decay by adding its reciprocal to that of the bulk
recombination time. It was shown [28–30] that surface recombination can
be as well presented in a more intuitive way through the surface recombi-
nation velocity (cm/s),

L
s= (11)
2[τ s − L (π 2 D 2 )]
2
26 Physics of Thin-Film Photovoltaics

where L is the device thickness, and D is the carrier diffusion coefficient.


As an illustration, we mention the fitting results from Ref. [3] related to the
above mentioned asymmetric p-n junction parameters, sn = 3 × 104 cm/s
and sp = 100 cm/s and demonstrating again strong asymmetry between
highly defective n- and more perfect p-materials. Some of the existing pho-
tovoltaic technologies use surface passivation (with Hydrogen, sulphur, or
other suitable materials) to suppress the corresponding recombination
velocities.
The current voltage characteristic of Eq. (6) can be readily derived for
the case of p-n diode by relating p- and n-type layers to the metals with
electrodes having Fermi levels EF2 and EF1 in Fig. 9. However, the satura-
tion current density is presented now as [12]

1 LDn 1 LDp G
J0 eN c N v exp . (12)
Nd n Na p kT

The activation energy G (instead of the Fermi energy) reflects the


nonequilibrium nature of the process where thermally generated charge
carriers are swept away of the system (not enough time for the electron
thermalization down to Fermi energy).
A historically significant advantage of the Fig. 15 p-n model is its suit-
ability for analyses utilizing the drift-diffusion kinetics, defect facilitated
recombination processes, and light absorption. That made it possible to
relate the device performance with its constituting material and structure
parameters and provided a solid guide for crystalline PV developments.
Here, we will not reproduce the corresponding cumbersome analytical
results found with many standard textbooks [2−12] and not practically
important nowadays due to the availability of simple modeling software
solving 1D diode problems.

D. Heterojunctions
Two dissimilar semiconductors form a heterojunction. Its band diagram is
obtained by the same general rule that Fermi levels in the component mate-
rials must be equal. It is always true that the Fermi level is uniquely related
to the work function, which is the energy difference between the electron
at the vacuum and Fermi levels. However, in semiconductors, the Fermi
energy depends on doping and thus is not a material parameter. Instead,
band edges should be considered as the parameters uniquely related to a
semiconductor. The position of conduction band edge relative to vacuum
One-Dimensional (1D) Diodes and PV 27

level is described by the electronic affinity, defined as the energy obtained


by moving an electron from the vacuum just outside the semiconductor
to the bottom of the conduction band just inside the semiconductor. The
electronic affinities are generally different for different semiconductors.
Because of that difference matching band diagrams for heterojunctions can
exhibit discontinuities (conductive band offsets) illustrated in Fig. 16 of the
type that depends on the electron affinities involved.
Shown in Fig. 17 is a commonly used (‘baseline’) band diagram of
CdTe or CIGS based thin film PV. Most of researchers tend to interpret
it in terms of p-n heterojunction assuming p-type CdTe (or CIGS) and
n-type CdS. That interpretation dominates even though there is no exper-
imental proof for n-type CdS. Furthermore, the existing data show that
CdS performs best when prepared highly resistive, thin, and significantly
non-crystalline of rather p-type. The miraculous CdS layer appears hardly
replaceable with significantly different materials, and a possible resolution

CBO>0 G2 CBO=0 CBO<0


G2
energy

G2
G1 G1 G1

Fig. 16 Possible band diagram alignments with positive, zero, and negative Conductive
Band Offset (CBO) corresponding to the cliff, smooth, and barrier type of band structure.

CdTe or CIGS

+
TCO
+
+
CdS Back
buffer
metal

Fig. 17 A rough sketch of the commonly used band diagram for CdTe and CIGS based
thin film PV. The CdS forbidden gap (≈ 2.4 eV) is greater than that of CdTe (≈ 1.5 eV)
or CIGS (≈ 1 − 1.7 eV depending on composition). Such combinations of different gap
materials are called heterojunctions. The dash and dot depicted features in the CdS region
represent a hypothetical energy barrier or a cliff introduced by some researchers to better
fit data; they can be related to the band structure mismatches.
28 Physics of Thin-Film Photovoltaics

of that mystery can be related to the piezo-photovoltaic coupling that


becomes possible due to the well-known (mostly beyond the PV com-
munity) piezo- and pyro-electric properties of CdS in combination with
the compressive stress created by the deposition and postdeposition pro-
cesses. While these issues will be addressed in what follows (Sec. XVIII)
we note here that all the ‘successful’ replacement of CdS layer (such as e.g.
MgZnO) utilized materials that are piezo-active as well.
The same figure presents other barrier features. On the back, there is a
Schottky barrier originating from the difference between work functions
of CdTe and the metal electrode. The front part can accommodate either
an energy cliff or energy barrier (according to different authors) reflecting
the data used for determining the band structure mismatch between CdS
and front electrode or buffer layer. More sophisticated diagrams presenting
this or other specific device recipes can include other barriers. However,
regardless of all these complementary details, the mainstream understand-
ing of such devices is that of p-n junction.

E. Other Relevant Types of Diodes


p-i-n barriers presented in Fig. 18 have been historically important with
a-Si:H thin film PV technology [31]. They consist of two heavily doped p-
and n- layers separated by the undoped intrinsic region. In the latter, both
the screening lengths of Eqs. (7) and (8) exceed the layer thickness leaving
the field in it practically uniform, which provides efficient drift of photo-
generated charge carriers towards the electrodes. In reality, such a diagram
may not be limited to a-Si:H. It may approximately apply to CdTe based
devices where screening length is of the same order as the layer thickness.
We shall end this section by noting that the current voltage characteristic
of Eq. (6) can be readily extended over all the above mentioned diodes by
relating the junction forming layers to the metal electrodes having Fermi
levels EF2 and EF1 in Fig. 9. The expression for saturation current density is

p intrinsic n

EF

Fig. 18 The schematic structure and band diagram of a p-i-n diode.


One-Dimensional (1D) Diodes and PV 29

not immediately apparent and remains a model parameter extracted from


the data or further theoretical estimates.
It is important that any of the above discussed or other conceivable
diodes can be described in detail based on the established equations, which
are the electrostatic Poisson equation and the continuity equation includ-
ing the recombination processes. The corresponding analytical treatments
can be rather involved. However there exists various software packages
that generate all the required characteristics given the input of material
parameters. They will be introduced in Sec. III B below.

F. Field Reversal Diode: A Counterintuitive Case


The versatility of diode concept exhibits itself with a rather counterintui-
tive arrangement where the electric field undergoes sign reversal includ-
ing its discontinuity, and the electric potential distribution shows a gull
wing singularity shape as illustrated in Fig. 19. Here we will not provide
its analytical and numerical analysis referring to Sec. XVIII below where
it is presented for a more general case including photocurrent. We would
like to note however that the field reversal diode can be quite realistic
corresponding to the case where one of the components (A in Fig. 19)
is ferroelectric, i.e. it has structural electric dipoles that can align to the
electric field.
Indeed, aligning its dipoles in such direction that their negative
poles point
  right in Fig. 19 will decrease the system electrostatic energy
Wel p E (another wording of the same would be that the positive
(­negative) poles move along (against) the field decreasing the system
Electron energy
Field
Electron energy

A B
Field

A B

Fig. 19 The electron energy and field distribution for a two-layer (A,B) system where
one of the components is ferroelectric, capable of aligning its dipoles to the external field.
Left: the standard distribution where the left side layer has n- and the right-side layer
p - type of conductivity and the built-in filed is significant in the junction proximity.
Right: the distributions after the electric dipoles in the ferroelectric layer (A) are aligned
minimizing the system energy. Their created electric field is opposite to the original one
leading to field reversal. The field discontinuity at the material contact plane is due to the
polarization surface charge density.
Another random document with
no related content on Scribd:
back
back
back
back
back
back
back
back
back
back

You might also like