Sunteți pe pagina 1din 786

FUNDAMENTALS OF

SEMICONDUCTOR PHYSICS AND DEVICES


Rolf Enderlein
ihmboldt-bhiuemity BerEin
Universi?yof Sao Pauh

NommJ.M. Horing
stewwas Institute of Technology
Hoboken, NJ

World Scientific
Singapore New Jersey London Hong Kong
Puhli.dzed hy
World Scientific Publishing Co. Re. Ltd.
P 0 Box 128, Farrer Road, Singapore 912805
USA oflice: Suite 1 H, 1060 Main Streei, River Edge, NJ 07661
UK oflcficer 57 Shelton Street, Covcnt Garden, London WCZH 9IiE

British Library Catalogiiing-in-PublicationData


A catalogue recurd fur this book is available from the British Library.

First published 1997


Reprinted 1999

FUNDAMENTALS OF SEMICONDIJC'IOK PHYSICS AND DEVICES


Copyright 0 1997 by World Scicntific Publishing Co Pte. Ltd.
All rights reserved. Th.is book, or parts thrreof, may ~ O be
I reprudrtced in any jurwi or by ony m e w s ,
electmnir or rnerhirnicirl, incIudinx photocopying. recording or any information storage and retrieval
sys:slew now known or m be ii-zvmted, without written permissionfrom !he Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive. Danvers, MA 01923, U S A . In this casc pcmission to
photocopy i s not required from the publisher.

ISBN 981-02-2387-0

Printed in Singapore.
Dedication
This book is dedicated to the memory of Adele and Werner Enderlein (par
ents of R.E.), and to t h e memory of Joseph and Esther Morgenstern (hfor
ganstein)(grandparents of N.J. M. 11.).
Vii

Preface

People come to technical books with a vast array of daerent needs and
requirements, arising from their differing educational backgrounds, profes-
sional orientations and career objectives. This is particularly evident in the
field of semiconductors, which stands at the juncture of physics, chemistry,
electronic engineering, material science and mathematics. No longer just
an academic discipline. this field is at the heart of an ongoing revolution in
communications, computation and electronic device applications, that inno-
vate many fields and change modern life in myriad ways, large and small. Its
profound impact and further potential command interest and attention from
all corners of the earth. and from a wide variety of students and researchers.
The clear need to address a broadly diversified and variously motivated
readership has weighed heavily in the authors considerations. It poses a
pedagogical problem faced by many teachers of intermediate level courses
on semiconductor physics. Generally speaking, every student has previously
studied about half of the course materiaL The difficulty lies in the fact that
each students exposure is likely to have been to a &fleerent half, depend-
ing on which lower level courses and teachers they have had, and where the
emphasis lay. To accommodate readers with varied backgrounds we start
from first principles and provide fully detaiIed explanations and proofs, as-
suming only that the reader is familiar with the Schrodinger equation. This
intensively tutorial treatment of the electronic properties of semiconduc-
tors includes recent fundamental developments and is carried through to the
physical principles of device operation, to meet the needs of readers inter-
ested in engineering aspects of semiconductors, as well as those interested in
basic physics. Clarity of explanation and breadth of exposure relating to the
electronic properties of semiconductors, from first principles to modern de-
vices, are our principal objectives in this fraddy pedagogical book. We offer
full mathematical derivations to strengthen understanding and discuss the
physical significance of results. avoiding reliance on hand waving arguments
alone.
To support the readers introduction to the physics of semiconductors, we
provide a thorough grounding in the basic principles of solid state physics,
assuming no prior knowledge of the field on the part of the reader. An ele
mentary discussion of the crystal structure, chemical nature and macroscopic
properties characterizing semiconductors is given in Chapter 1. Moreover,
we also include an extensive appendix to guide the reader through group
theory and its applications in connection with the symmetry properties of
semiconductors, which are of major importance. Beside spatially homoge-
neous bulk semiconductors, we undertake a full exposition of inhomogeneous
semiconductor junctions and heterostructures because of their crucial role
Preface iu

The book has emerged from lectures which the authors presented for physics
students at the Humboldt-University of Berlin. Germany, and the State Uni-
versity of Sao Paulo, Brazil, and for physics and engineering physics students
at the Stevens Institute of TeFhnology in Hoboken, New Jersey, C.S.A. Part
of the book has similarities with the german book "Grundlagen der Halbleit-
erphysik" ("Fundamentals of Semiconductor Physics") which was written by
one of us (R. E.) together with A. Schenk. We are thankful to Dr. Schenk
(now at ETH Zurich) for allowing us to use part of his work in the present
volume. In writing this book we have had excellent suppoIt from many
of our colleagues at our own and other Universities. We are particularly
thankful to Prof. Dr. J. Auth (Humboldt-Lniversity Berlin), Prof. Dr. F.
Bechstedt (Friedrich-Schiller University Jena), Prof. Dr. W. A. Harrison
(Stanford University), Prof. Dr. M. Scheffler (Fritz-Haber Institut, Berlin),
Prof. Dr. J. R. b i t e , Prof. Dr. A. Fazzio, and Prof. Dr. J. L. Alves
(State University Sao Paulo), as well as to Prof. Dr. H. L. Cui, Prof. Dr.
G. Rothberg, Mr. G. Lichtner (Stevens Institute of Technology), and Prof.
Dr. G . Gumbs (Hunter College, CWNY, New York), who read parts of the
manuscript and contributed helpful suggestions and critical remarks. The
technical assistance of Mrs. Hannelore Enderlein is gratefully acknowledged.

RoIf Enderlein Norman J.M. Horing


Sao Paulo Hoboken, N J
October 1996
Xi

Contents
1 Characterization of sernicond uct ors 1
1.1 Inlrnduclion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Atomic structure of ideal crystals . . . . . . . . . . . . . . . . 5
1.2.1 Cryst.al latlices . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 Point groups of equivalent directions arid crystal classes 12
1.2.3 Space groups and crystal structures . . . . . . . . . . 14
1.2.4 Cubic semiconductor structures . . . . . . . . . . . . . 16
1.2.5 Hexagoiial semiconductor st.ructures . . . . . . . . . . 22
1.3 Chemical nature of semiconductors. Material classes . . . . . 28
1.3.1 Group IV elemental semiconductors . . . . . . . . . . 29
1.3.2 111-V semiconductors . . . . . . . . . . . . . . . . . . . 30
1.3.3 11-VI semiconductors . . . . . . . . . . . . . . . . . . . 31
1.3.4 Group \'I elemental semiconductors . . . . . . . . . . 31
1.3.5 IV-VI semiconductors . . . . . . . . . . . . . . . . . . 32
1.3.6 Other compound semiconductors . . . . . . . . . . . . 32
1.4 hlacroscopic properties and their microscopic implications . . 33
1.4.1 Electrical conductivity . . . . . . . . . . . . . . . . . . 34
1.4.2 Depenclenre of conductivity on the semiconductor state 35
1.4.3 Optical absorption spectrum and the band modcl of
srmicoiiductors . . . . . . . . . . . . . . . . . . . . . . 38
1.4.4 Electrical conductivity in the band model . . . . . . . 43
1.4.5 The Hall effect and the existence of positively charged
freely mobile carriers . . . . . . . . . . . . . . . . . . . 45
1.4.6 Seinicondiictors far from thermodynamic equilibrium . 49

2 Electronic structure of ideal crystals 51


2.1 Abcimic cores and vdcnce electrons . . . . . . . . . . . . . . . 51
2.2 The ciynaniical problem . . . . . . . . . . . . . . . . . . . . . 54
2.2.1 Schriidiiiger equation for the interacting core and va-
lence dwtl-on system . . . . . . . . . . . . . . . . . . . 54
2.2.2 Adiabatic approximation . Lattice dynamics . . . . . . 57
2.2.3 Oneparticle approximation . Oneparticle Schriidinger
equation . . . . . . . . . . . . . . . . . . . . . . . . . . 66
2.3 General properties of stationary one-rlectron states in a crystal 82
2,3.1 Syinrnctry properties of the one-electron Schrtidinger
equation . . . . . . . . . . . . . . . . . . . . . . . . . . 82
2.3.2 R b c h theorem . . . . . . . . . . . . . . . . . . . . . . 85
2.3.3 Reciprocal v e c h space and the reciprocal latt.ice . . . 89
2.3.4 Relation between energy eigenvalues and quasi-wave
vector . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
xii Contents

2.4 Schrodinger equation solution in the nearly-freeelectron ap-


proximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.4.1 Kon-degenerate perturbat.ion t.heory . . . . . . . . . . 100
2.4.2 Degenerate perturbation theory . . . . . . . . . . . . . 103
2.5 Bandstructure . . . . . . . . . . . . . . . . . . . . . . . . . . 105
2.5.1 Brillouin zones . . . . . . . . . . . . . . . . . . . . . . 105
2.5.2 Degeneracy of energy bands . . . . . . . . . . . . . . . 116
2.5.3 Critical points and effective masses . . . . . . . . . . . 119
2.5.4 Density of states . . . . . . . . . . . . . . . . . . . . . 123
2.5.5 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2.5.6 Calculational methods for band structure determination133
2.6 Tight binding approximation . . . . . . . . . . . . . . . . . . 140
2.6.1 Fundamentals . . . . . . . . . . . . . . . . . . . . . . . 140
2.6.2 TB theory- of diamond and zincblende type semicon-
ductors . . . . . . . . . . . . . . . . . . . . . . . . . . 148
2.6.3 sp3-hybrids, total energy and chemical bonding . . . . 165
2.7 k . p-met.hod . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
2.7.1 Fundamentals . . . . . . . . . . . . . . . . . . . . . . . 179
2.7.2 Valence bands of diamond structure semiconductors
without spin-orbit interaction . . . . . . . . . . . . . 184 .
2.7.3 h t t i n g eT-Kohn model . . . . . . . . . . . . . . . . . . 189
2.7.4 Kana model . . . . . . . . . . . . . . . . . . . . . . . . 200
2.8 Band structure of important semiconductors . . . . . . . . . . 211
2.8.1 Silicou . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
2.8.2 Germanium . . . . . . . . . . . . . . . . . . . . . . . . 218
2.8.3 111-V semiconductors . . . . . . . . . . . . . . . . . . 219
2.8.4 IGVI semiconductors . . . . . . . . . . . . . . . . . . . 221
2.8.5 IV-\'I semiconductors . . . . . . . . . . . . . . . . . . 224
2.8.6 Tellurium and selenium . . . . . . . . . . . . . . . . . 224

3 Electronic s t r u c t u r e of semiconductor crystals with p e r t u r -


bations 225
3.1 Atomic structure of red semiconductor crystals . . . . . . . . 226
3.1.1 Classification of perturbations . . . . . . . . . . . . . . 226
3.1.2 Point perturbations . . . . . . . . . . . . . . . . . . . . 227
3.1.3 Formation of point perturbations and their movenient 235
3.1.4 h e and planar defects . . . . . . . . . . . . . . . . . 240
3.2 One-electron Schrodinger equation for point perturbations . . 241
3.2.1 Electron-core interaction . . . . . . . . . . . . . . . . . 242
3.2.2 Electron-elw?c.lroninteraction . . . . . . . . . . . . . . 245
3.3 Effective mass equation . . . . . . . . . . . . . . . . . . . . . 252
3.3.1 Effectivemass equation for a single band . . . . . . . 253
3.3.2 Multjband effective mass equation . . . . . . . . . . . 259
Contents
...
Xlll

3.4 Shallow levels. Donor and acceptor states . . . . . . . . . . . 265


3.4.1 Hydrogen model . . . . . . . . . . . . . . . . . . . . . 266
3.4.2 Improvements upon the hydrogen model . . . . . . . . 272
3.5 Deeplevds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
3.5.1 General characterization of deep levels . . . . . . . . . 281
3.5.2 Defect molecule model . . . . . . . . . . . . . . . . . .285
3.5.3 Solution methods for the oneelectron Schriidinger q u a -
tion of a crystal with a point perturbation . . . . . . . 293
3.5.4 Correlation effects . . . . . . . . . . . . . . . . . . . . 301
3.5.5 Resu1t.s for se1ecDed deep centtas . . . . . . . . . . . . 308
3.6 Clean semiconductor surfaces . . . . . . . . . . . . . . . . . . 334
3.6.1 The concept of clean surfaces . . . . . . . . . . . . . . 334
3.6.2 Atomic structure of clean surlaces . . . . . . . . . . . 336
3.6.3 Electronic structure of crystals with a surface . . . . . 354
3.6.4 htomic and electronic structure of particular surfaces 371
3.7 Semiconductor microstructures . . . . . . . . . . . . . . . . .388
3.7.1 Neterojunctions . . . . . . . . . . . . . . . . . . . . . . 388
3.7.2 Microstructures; Fabrication, classifications, examples 396
3.7.3 h*lethodsfor electronic structure calculations . . . . . 409
3.7.4 Elcctronic structure of particular microstructures . . . 420
3.8 Macroscopic electric fields . . . . . . . . . . . . . . . . . . . . 433
3.8.1 Effective mass equation and stationary electron states 434
3.8.2 Non-stationary states . Bloch oscillations . . . . . . . . 437
3.8.3 Interband tunneling . . . . . . . . . . . . . . . . . . . 440
3.8.4 Photon assisted interband tunneling . . . . . . . . . . 442
3.9 Macroscopic magnetic fields . . . . . . . . . . . . . . . . . . . 443
3.9.1 Effective mass equation in a magnetic field . . . . . . 444
3.9.2 Solution of the effective mass equation . . . . . . . . . 452
4 Electron system in t herrnodynamic equilibrium 457
4.1 Fundamentals of the statistical description . . . . . . . . . . . 457
4.2 Calculation of average particle numbers . . . . . . . . . . . . 460
4.2.1 Configuration-independent oneparticle states . . . . . 460
4.2.2 Configuration-dependent one-particle states . . . . . . 462
4.3 Density of states . . . . . . . . . . . . . . . . . . . . . . . . . 469
4.3.1 Total electron concentration . . . . . . . . . . . . . . . 469
4.3.2 Density of states of ideal semiconductors . . . . . . . . 470
4.3.3 Density of states of real semiconductors . . . . . . . . 474
4.4 Free carrier concentrations . . . . . . . . . . . . . . . . . . . . 477
4.4.1 Conservation of total electron number . . . . . . . . . 477
4.4.2 Free carrier concentration dependence on Fermi en-
ergy. Law of mass action . . . . . . . . . . . . . . . . . 478
4.4.3 Intrinsic semiconductors . . . . . . . . . . . . . . . . . 482
xiv Contents

4.4.4 Extrinsic semiconductors . . . . . . . . . . . . . . . . 484


4.4.5 Compensation of donors and acceptors . . . . . . . . . 489
4.4.6 More complex cases . . . . . . . . . . . . . . . . . . . 492

5 Non-equilibrium processes in semiconductors 499


5.1 Fundamentals of the statistical description of non-equilibrium
processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
5.2 Systematics of non-equilibrium processes in semiconductors . 505
5.2.1 Temporal inhomogeneity and spatial homogeneity . . 505
5.2.2 Spatial inhomogeneity and temporal homogeneity . . . 506
5.2.3 Space and time inhomogeneities . . . . . . . . . . . . 508
5.3 Generation and annihilation of free charge carriers . . . . . . 509
5.3.1 Generation processes . . . . . . . . . . . . . . . . . . . 510
5.3.2 Unipolar annihilation of free charge carriers: capture
at deep centers . . . . . . . . . . . . . . . . . . . . . . 511
5.3.3 Bipolar annihilation of carriers at deep centers . . . . 517
5.4 Drift current . . . . . . . . . . . . . . . . . . . . . . . . . . . 523
5.5 Diffusion and annihilation of free carriers . . . . . . . . . . . 527
5.6 Equilibrium of free carriers in inhomogeneously doped semi-
conductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530

6 Semiconductor junctions in thermodynamic equilibrium 535


6.1 pn-junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537
6.1.1 Establishment of thermodynamic equilibrium . . . . . 539
6.1.2 Diffusion voltage . . . . . . . . . . . . . . . . . . . . . 541
6.1.3 Spatial variation of the electric and chemical poten-
tials: Schottky approximation . . . . . . . . . . . . . . 542
6.2 Heterojunctions . . . . . . . . . . . . . . . . . . . . . . . . . . 549
6.2.1 Equilibrium condition . . . . . . . . . . . . . . . . . . 550
6.2.2 Electrostatic potentid . GaAs/Gal-,Al, As heterojunc-
tion as an example . . . . . . . . . . . . . . . . . . . . 552
6.3 Metal-semiconductor junctions . . . . . . . . . . . . . . . . . 557
6.3.1 Energy level diagram before establishing equilibrium . 357
6.3.2 Electrostatic potential . . . . . . . . . . . . . . . . . . 559
6.3.3 Schottky barrier . . . . . . . . . . . . . . . . . . . . . 563
6.4 Insulator-semiconductor junctions . . . . . . . . . . . . . . . . 567
6.4.1 Thermodynamic equilibrium . . . . . . . . . . . . . . 567
6.4.2 Influence of interface states . . . . . . . . . . . . . . . 570
6.4.3 Semiconductor surfaces . . . . . . . . . . . . . . . . . 572

7 Semiconductor junctions under non-equilibrium conditions 573


7.1 pn-junction in an external voltage . . . . . . . . . . . . . . . 574
7.1.1 Electrostatic potential profile . . . . . . . . . . . . . . 576
Contents xv

7.1.2
Mechamism of current transport through a pn-junction 577
Chemical potential profiles for electrons and holm . . 580
7.1.3
Dependence of current density on voltage . . . . . . . 583
7.1.4
Bipolar transistor'. . . . . . . . . . . . . . . . . . . . .
7.1.5 585
7.1.6 T u n e 1 diode . . . . . . . . . . . . . . . . . . . . . . . 593
7.2 yn-junction in interaction with light . . . . . . . . . . . . . . 595
7.2.1 Photocffect at a pn-junction . Photodiode and photo-
voltaic element . . . . . . . . . . . . . . . . . . . . . . 595
7.2.2 Laser diode . . . . . . . . . . . . . . . . . . . . . . . . 599
7.3 Metal-semiconductor junction in an external voltage.
Rectificrs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 606
7.4 hwulator-semiconductor junction in an external voltage . . . 612
7.4.1 Field effect . . . . . . . . . . . . . . . . . . . . . . . . 612
7.4.2 Inversion layers . . . . . . . . . . . . . . . . . . . . . . 614
7.4.3 MOSFET . . . . . . . . . . . . . . . . . . . . . . . . . 620

Appendices
A Group theory for applications in semiconductor physics 623
A.1 Definitions and concepts . . . . . . . . . . . . . . . . . . . . . 623
A . l .1 Group definition . . . . . . . . . . . . . . . . . . . . . 623
A.1.2 Concepts . . . . . . . . . . . . . . . . . . . . . . . . . 624
A.2 Rigid displacements . . . . . . . . . . . . . . . . . . . . . . . 627
A.2.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . 621
A.2.2 Translations . . . . . . . . . . . . . . . . . . . . . . . . 628
A.2.3 Orthogonal transformations . . . . . . . . . . . . . . . 629
A.2.4 Geometrical interpretation . . . . . . . . . . . . . . . . 631
-4.2.5 Screw rotations and glide re3ections . . . . . . . . . . 632
A.3 Translation. point and space groups . . . . . . . . . . . . . . 635
A.3.1 Lattice translation groups . . . . . . . . . . . . . . . . 635
-4.3.2 Point groups . . . . . . . . . . . . . . . . . . . . . . . 636
A.3.3 Space groups . . . . . . . . . . . . . . . . . . . . . . . 654
A.4 Representations of groups . . . . . . . . . . . . . . . . . . . . 655
A.4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . 655
A.4.2 Irreducible representations . . . . . . . . . . . . . . . . 661
4.4.3 Products of representations . . . . . . . . . . . . . . . 667
A . 5 Representations of the full rotation group . . . . . . . . . . . 673
4.5.1 Vector representation of the rotation group and gen-
erators of infinitesimal rotations . . . . . . . . . . . . 674
A.5.2 Representations for dimensions other than three . . . 676
A.6 Spinor representations . . . . . . . . . . . . . . . . . . . . . . 682
A.6.1 Space-dependent spinors . . . . . . . . . . . . . . . . . 682
mi Cootents

A.6.2 Representation V I . . . . . . . . . . . . . . . . . . . . 683


2
A.6.3 Irreducible spinor representations . . . . . . . . . . . . 684
A.6.4 Double group method . . . . . . . . . . . . . . . . . . 685
A.7 Projective representations . . . . . . . . . . . . . . . . . . . . 687
A.7.1 Factor systems . . . . . . . . . . . . . . . . . . . . . . 687
A.7.2 Definitions and theorems . . . . . . . . . . . . . . . . 689
A.7.3 Construction of projective representations . . . . . . . 692
A.8 Time reversal symmetry . . . . . . . . . . . . . . . . . . . . . 692
A.8.1 Time reversal operator . . . . . . . . . . . . . . . . . . 693
A.8.2 Additional degenerac!?~of energy eigenvalues . . . . . 694
A.8.3 Additional selection rules for matrix elements . . . . . 697
A.9 Irreducible representations of space groups . . . . . . . . . . . 698
A.9.1 Representations of translation groups . . . . . . . . . 698
A.9.2 Star of wavevectors . . . . . . . . . . . . . . . . . . . . 700
A.9.3 Small point groups and their projective representations 702
A.9.4 Representations of the fufl space group . . . . . . . . . 704
A.9.5 Spinor representations of space groups . . . . . . . . . 706
A.9.6 Implications of time reversal symmetry . . . . . . . . 707
A.9.7 Compatibility . . . . . . . . . . . . . . . . . . . . . . . 712
A.10 Irreducible representations of small point groups . . . . . . . 712
A.10.1 Character tables . . . . . . . . . . . . . . . . . . . . . 712
A.10.2 Multiplication tables . . . . . . . . . . . . . . . . . . 731
A.10.3 Compatibility relations . . . . . . . . . . . . . . . . . 734

B Corrections to the adiabatic approximation 737

C Occupation number representation '741

Bibliography 747

Index 757
FUNDAMENTALS OF
SEMICONDUCTOR PHYSICS AND DEVICES
1

Chapter 1

Characterization of
semiconductors

1.1 Introduction
Semiconductors are identified as a unique material group on the basis of
their common macroscopir properties, as is done for metals, dielectrics and
magnetic materials. The name semiconductor stems from the fact that
such materials have moderately good conductivity, higher than that of in-
sulators, and lower than that of metals. However, if this were the only
property which these materials had in common, the term semiconductor
would have only a very weak foundation. But such is not the case. In fact,
many materials having conductivity between that of metals and insulators.
display simultaneously a series of further common properties. In particular,
their conductivity depends very strongly on material staie, for example, on
temperature and chemical purity, much more so than in the case of met-
als. For sufficiently pure semiconductors, the conductivity decays by orders
of magnitude while cooling down from room temperature to liquid helium
temperature. ,4t absolute zero temperature, semiconductor conductivity al-
most vanishes, in contrast to the conductivity of metals, which rises modestly
with falling temperature. The conductivity of metals reaches its maximum
at low temperature, and for superconductors it effectively becomes infinitely
large. In regard to the dependence of semiconductor conductivity on the
degree of purity, it has been found that a given semiconductor in a very
pure state can resemble an insulator. while in a highly polluted state it acts
like a metal, among other peculiarities. Furthermore, irradiation with light
can cause a transition from insulator-like behavior to metal-like behavior of
one and the same semiconductor.
There are yet other optical properties shared by semiconductors: The op-
2 Chapter I. Characterization of semiconductors

tical absorption spectra of semiconductors exhibit a threshold - below the


threshold frequency, light can pass through practically without losses, while
above it the light is strongly absorbed. Moreover, good luminescence proper-
ties in the visible and infrared spectral range are also characceristic of many
semiconductors. Thus, the identification of a semiconductor material in-
volves several characteristic properties, not just the one of moderately good
electrical conduction. One has a semiconductor only if all such properties
apply. This criterion excludes ionic conductors, for example, which exhibit
conductivity values of the right order of magnitude, but do not display the
characteristic temperature dependence.
One may justifiably question the extent to which this definition of a semi-
conductor really makes sense. For example, it is not yet obvious why just
the above properties are selected as defining features while others are not,
and what internal connections may exist among them. A full answer can
only be given by means of a microscopic theory of semiconductor properties
which will be developed systematically in the chapters to follow. For the
moment, we invite the reader to join us in the recognition that all macro-
scopic properties involved in the definition of a semiconductor can be traced
back to a common microscopic origin, namely the nature of the spectrum
of allowed energy levels and the particulars of their population by electrons.
To be specific, the permissible energy levels of a semiconductor form bands
which are separated by forbidden regions, and the Characteristic electron
population of allowed energy levels is such that, at absolute zero tempera-
ture, a semiconductor i s characterized by having only completely occupied
and completely empty energy bands (no partially filled bands). It is this
common microscopic feature which underlies the totality of macroscopic ma-
terial properties that uniquely define a semiconductor. It also provides the
basis for uncovering yet other common macroscopic features of this class
of materials, beyond the ones already discussed. For instance, it may be
expected that semiconductors should be predominantly solid crystalline ma-
terials, since the formation of energy bands with gaps between them is most
likely to occur in the crystalline phase. Nevertheless, amorphous and liq-
uid semiconductors cannot be completely rxcluded since a certain regularity
of the relative positions of neighboring atoms also exists in the amorphous
and liquid phases. Actually, in addition to solid crystalline semiconductors,
which are the main reprcsentatives, a series of amorphous semiconductors
has also been found silicon and selenium being important examples. Liq-
uid semiconductors are also possible, with melted tellurium among them.
Other semiconducting materials, e.g., silicon and germanium, are metals in
the liquid phase.
In this book we restrict our considerations to solid crystalline semicon-
ductors. The discussion also partially applies to amorphous and liquid sexni-
conductors, but in most cases modifications are necessary. Even the basic
1.1. Introduction 3

concept o f the quantum mechanical energy spectrum uf electrons has to


be defined in tl different way, and a proper treatment of amorphous and
liquid materials cannot be acconimodated within the framework of this in-
troduction to semiconductor physics. Interested readers are referred to the
extensive literature on this subject (see, for example? Elliot, 1983; Mott
and Davis, 11979; Bunch-Bruevich, Enderlein, Ewer, Keiper, Mironov, and
Xvyagin, 1984; Adhr, F r h s c h e and Ovahinsky, 1985).
T h e microscopic definition discussed above contains no recognizable con-
straints with regard to the chemical nature of semiconductors. One may
expect, therefore, that semiconductors should be distinguished by having a
large chemical diversity. This is in fact true semiconductors may be com-
~

posed of a large variety of chemical elements and compounds. The most


fully expAored candidates and those used for technical applications today arc
crystalline semiconductors consisting of relatively small chemical u n i t s , i.e.
either dements or binary and ternary compounds.
Knowledge of thc mxistence of a distinct material group semiconductors
developed, historically, only relatively late. MPtals have been used by men
since antiquity, but semiconductors attracted attention for the Erst time only
a century and a half ago. The first reference to a characteristic semiconductor
property dates back to Faraday who in 1833 observed an increase of the
electric conductivity of silver sulfide with temperature. The exponential form
of this increase was discovcrcxl by Hittorf in 1851. The trerm semiconductor
was introduced in 1911 by Konigsberger and Weiss after a similar term of
about the same context had already been employed by Ebert (1789) and
Bromme (1851).
The late and, initially, relatively slow development of semiconductor
physics, is primarily due to the circnnistance that the characteristic proper-
ties of semiconductors depend strongly on their degree of purity, more pre-
cisely, on the presence 01absence of certain chemical elements. This is also
the reason that many semiconducting materials in their natural form as min-
erals do not display the typical properties of a semiconductor they are too
~

heavily polluted and have too many structural defects. Natural diamondrr,
for example, are semiconductors only in rare cases. Accordingly, clean fab-
rication of the materials in the laboratory and the controlled incorporation
of chemical elements played a crucial role from the very beginning. The lack
of such control in preparation presented fundamental difficulties which had
to be overcame in the early days of semiconductor rasearch. The necessary
accuracy of composition control, which amounts to one atom in one hundred
thousand or less, excccdcd the accuracy that prevailed in chemistry at the
t h e by orders of magnitude. It was necessary to raise the accuracy of chem-
ical composition control to a level wherein one could measure one millionth
of a mole haction instead of one thousandth. only in this way could reliable
results be achieved with semiconductors. Since such accuracy was achieved
4 Chapter 1. Characterization of semiconductors

only gradually, semiconductor physics, in its early history, was confronted


with apparently mysterious phenomena and contradictory results. For ex-
ample, silicon and germanium were first thought to be metals, until the
recognition that the impurity concentrations of certain elements were too
large to achieve semiconducting properties.
From the very beginning semiconductor physics research was impeded
by the need for expensive fabrication of material samples. The Geld could
develop only to the extent that good samples could be made available. Nat-
urally, the first samples weie either materials which could b r fabricated rela-
tively cheaply or materials which occurred in nature in a suitable form, albeit
not ideal. Among these first materials were metal sulfides, which like lead
sulfide in its mineral form as Galena and copper oxide (CuzO) and selenium
in their artificially grown form, displayed good semiconducting properties
even with relatively strong pollution and structural imperfections. In 1874,
Braun discovered that contacts between certain metal sulfides and metal tips
fxhibited different electrical resistance upon reversal of the polarity of the
applied voltage. Such point contact structures were used in radio receivers
as rectifiers at the beginning of our century. One can mark these point con-
tact rectifiers as the first semiconductor devices. Similar rectifying action
was also found for selenium and copper oxide. Moreover, a large change
of electrical conductivity could also be achieved in these materials through
irradiation by light. For selenium, this property was discovered as early as
1852 by Hittorf. Since the beginning of the 20-th century, this effect has also
been used practically in the selenium photocell. The first technical use of
copper oxide as a rectifier was accomplished in 1926 by Grondahl, followed
by rectifiers using selenium. The first practical application of copper oxide
in photocells was accomplished in 1932 by Lang.
Because of their technological importance, selenium and copper oxide
were the first semiconductors to be subjected to more detailed physical in-
vestigations. The semiconducting metal sulfides, selenides and tellurides
were already known earlier because of their good luminescence properties.
In the further exploration of these materials semiconducting behavior, lu-
minescence physics partially merged with semiconductor physics.
In the mid-nineteen-thirties, the search for a solid-state-based electronic
switching element which could replace the vacuum tube was extended to
the elemental semiconductors germanium and silicon. The most important
results of this research, which turned out to be decisive for the whole further
development of semiconductor physics, were the invention of the germanium-
based bipolar transistor in 1949, and the realization of the field effect tran-
sistor with the help of silicon at the end of the nineteen-fifties. With the
introduction of silicon, the stage was set for the development of semiconduc-
tor microelectronics. Later, a similar role was played by compounds involv-
ing elements of the third and fifth groups of the periodic table, like gallium
1.2. Atornic structure of ideal m y s t a h 5

arsenide or phosphide, making possible the development of semiconductor


optoelectronirs.
The broad technical application of its results distinguishes semiconduc-
tor physics now from its early days. It is well established that semicondnc-
tors arc exceptionally well-suited for necessary functions in electronics and
electrical engineering. This is by no means accidental, but is due to the mi-
croscopic nature of semiconductors, which permits the controlled variation
of characteristic matprial properties by external means over a wide range
of parameters. The great technical importance of semiconductors has made
thorough physical investigation of these materials necessary, but it has also
justified the high cost involved in their fabrication and study. Owing to
both of these aspects semiconductors are thc best explored and understood
materials of condensrd matter today. Moreover, a multitude of physical phe-
nomena which occur in other solid state materials may also be observed in
semiconductors, often in the most distinctive way. For this reason studies
of semiconductors can also provide knowledge of other solid state mattri-
als. Semiconductors have in fact become model systems for basic research
in condensed matter physics.
The presentation of the microscopic principles of semiconductor physics
will occupy most of this book. The introductory first chapter lies outside
of this framework because it involves discussion of the results on a phe-
nomenological basis. The characterization of semiconductors by means of
their unique macroscopic features, which we have touched upon above, will
be continued in Chapter 1. In this regard, atomic structure will be discuss4
in section 1.2, chemical nature in section 1.3 and macroscopic physical prop-
erties in section 1.4. In dealing with macroscopic properties we will not
restrict ourselves to mere description, but we will also use them to mo-
tivate the microscopic model of semiconductors introduced above. In this
connection, we will make the first step towards a microscopic theory of semi-
conductors in section 1.4. Naturally, this will have to be done in a heuristic
way, and many questions postponed until later. The full presentation of the
microscopic principles of semiconductor physics will follow in later chapters.

1.2 Atomic structure of ideal crystals


In all solid state materials, including amorphous ones, the neighbors of an
arbitrarily selected atom are ordered in a regular way, just as in a molecule.
The term short-range order is uscd for this property. The neighbors of a
particular atom form its short-range order complex. Semiconductor materi-
als, as they are treated in this book, are crystals. Sheir atomic structure is
approximately that of ideal crystals. The latter are distinguished by yet an-
other order, apart from short-range, which is termed long-range order. This
6 Chapter 1. Characterization of semiconductors

means that,, for a given atom, there are remote atoms possessing the same
short-range order complexes as the original atom, rand that the positions of
the remote atoms are related to the position of the origina1 atom by sirn-
ple transformations. 'l'hp crystal is considered t o be inhitely large in this
context. Atoms having identical short-range order complexes are termed
equivalent. Equivalent atoms are necessarily of the same chemical species,
but chemically identical atoms need not necessarily be equivalent.

1.2.1 Crystal lattices


The transformations displacing atoms into equivalent ones are translations
of the crystal by vectors T which form linear combiaations

of three non-coplanar vectors Al, Az. A3 with arbitrary integer coefficients


t l ,t 2 , t3. The parallelepipeds of the crystal spanned by the particular dis-
placement vectors Ak, A2, A3 are called unit cells. By putting unit cells
together the wholc crystal may be constructed. ' h e size of a unit cell and
the number of atoms in it is not fixed by the above definition, and can in
fact be taken arbitrarily large, as long 8s it remains finite. The pertinent
question is not how large a unit cell can be, but rather how small. The
answer to this question leads us t o the definition of the primitive unit cell
and the crystal lattice.

Definition

The smallest possible unit cell is called a primitive unit cell In the extreme
case this cell can contain only 1atom, but as a rule, it has several atoms. If
there is only one atom per cell, then the short-range order merges into the
long-range order.
If the unit cell is taken to be a primitive one, the vectors Al, A2, &
are some minimal vectors a l , az,w. The parallelepiped spanned by these
vectors is a primitive unit cell, Each translation by a vector R of the form

with integer coefficients r l , r2, r 3 transforms the crystal into itself. One refers
to this property as the trunslation~lsymmetry of the crystal. The point set
defined by the vectors R forms a spatial lattice called the crystal lattice. The
vectors al,a2, are termed primitive lattice vectors, The volume 00of a
primitive unit cell may be written as the triple scalar product of al,ag,a,

Ro = al . [ag x a].
1.2. Atomic structure of ideal cestals 7

While the lattice of a crystal and the volume Qo of its primitive unit cell
are well defined, this is not the case for its primitive lattice vectors al,a2,a3
and also it is not true for the form of its primitive unit cell. Any set of
linear combinations of the primitive vectors al,a2,a3 which yields a triple
scalar product equal to the volume Qo is again a set of primitive lattice
vectors, and the parallelepiped spanned by them forms a primitive unit cell.
The corners of the parallelepipeds do not necessarily have to lie on lattice
points. Each parallelepiped, shifted arbitrarily in space, is again a primitive
unit cell. The 'parallelepiped form is also not imperative, as there are also
other forms possible. An especially compact primitive unit cell is the so-
called Wzgner3eit.z cell. The center of this cell lies on a lattice point and its
surface is formed by the perpendicular bisector planes which divide in half
the line segments joining the center to adjacent lattice points.
Translations which transform a crystal into itself, by definition, do the
same for the lattice of the crystal. Here, the translations are through lattice
vectors R, called lattice tramtations The set of all lattice translations forms
a group. This term describes a mathematical set of elements among which
a 'multiplication' is defined that results in products which are also elements
of the set. Further properties of a set forming a group are listed in A p
pendix A. In particular, there must be an identity element, and the inverse
of an element must also be an element of the set. In the case of translations
the 'multiplication' is the consecutive application of two of these transfor-
mations. Since two consecutive lattice translations constitute yet another
lattice translation, and also the requirements of Appendix A are satisfied,
the set of all lattice translations of a crystal forms, in fact, a group, called
the translation symmetrg, group or simply the tramlation group. Groups of
symmetry elements play a central role in the description of the atomic struc-
ture and other microscopic properties of crystals. Appendix A provides a
thorough discussion of groups as needed in this book.

Point symmetry of lattices

We now ask whether there are other possible spatial transformations, besides
translations, which transform lattices into themselves. From the outset it is
clear that the only transformations which may be considered are those that
do not change the distances between lattice points, i.e. mgzd drsplacernents
of the lattice (see Appendix A). One can show that, besides translations,
there is a second class of transformations fitting this description, namely
rotatoom and reflectzons, as well as all products which are compounded from
them, such as rotation-reflections, rotatton-znverstonsand znverszon itself.
Taken together, they are termed orthogonal traasfomattom. These differ
from translations inasmuch as they leave one or several lattice points un-
changed, while the remaining points are shifted by vectors depending on
8 Chapter 1 . Characterkstion of semiconductors

their positions. In a t,ranslation, all points are shifted by the same vector,
with no points fixed. Rotations transform right-hand4 Cartesian coordi-
nate systems into right-handed systems, but the application of reflections
and inversion to right-handed system results in left-handed ones.
It turns out that there are orthogonal transformations which transform
lattices into themselves. They are called point symmetry operations of lat-
tices. The set of all point symmetry operations of a lattice forms a group,
as does the set of all lattice translations. The multiglicslion of two of these
operations is again understood to represent their consecutive application.
Groups of point symmetry operations are termed point gvoups. In Appendix
A we describe thwn in detail. Not all of the various point groups listed
in Appendix A are allowed as symmetry groups of crystal lailices, but only
parlicular ones which are called holohedral point groups. We will derive them
now by demonstrating that lhey must have three special properties.
First, all uf these point groups must contain the inversion transformation
with r e q e t t to the lattice point R = 0. This may be seen as fullows:
Inversion with respect to 0 transforms a lattice point R into -EL Considered
joiutly with R, the point -R is a lattice point having -q, -r2, -7-3 as integer
coefficients. Therefore, inversion with respect to 0 transforms an arbitrary
lattice into itself. It follows immediately that inversion with respect l o any
other lattice point will do h e same.
Second, it turns out that rotation symmetry axes Ohrough lattice points
can only be 2-, 3-, 4- and 6-fold while ri, 7- and more-fold axes are not
compatible with the translation symmetry of the lattice. One may prove
this as follows: Let C, be a rotation about such an axis t,hrough an angle
2?r/n. We consider a lattice plane perpendicular to this axis and denote a
primitive lattice vector of the corresponding planar lattice by f (see Figure
1.1). Rotating it through 2a/a, it becomes C,f, and a rotation by - 2 ? r / R
transforms it into Cqlf. If, as w e suppose, Cmbelongs to the point group of
the lattice, them s n d o a C;'. Thus both C,f and C i l f are vectors of the
plane lattice. The same holds for the sum Cnf i Cg'f ol t,he two vectors.
Moreover, C,,f + CGLf represents a vector parallel to f. This means that
C,f + CF'f must be an integer multiple of f. Since the largest possible
+
length of Cnf Cg'f can only he 2 / f J ,one has Cnf f C'L'f = p,F with
pn - -2. -LO, 1 or 2. This is indicative t.hat, the relation

p , = 2COS (?)
must hold for p,. For p n = -2, equation (1.4) yields n 2. For pn -= -1
1

one has n = 3, for pn - 0 it follows that n = 4 and for p , = 1 l h e solulion


is n = 6 . For ppI- 2 equation (1.4) has only the trivial solution n = I . This
completes the proof concerning rotation symmetv opexations. For so-caIled
gvasi-crystals. which do not exhibit an exact translation symmetry, rotations
1.2. Atomic structure of ideal crystals 9

Figure 1.1: On the mul-


tiplicity of the rotation
axes of crystal lattices.

about &fold and other axes are also possible.


Third, one finds that point groups of lattices having 3-, 4- or 6-fold axes
of rotation must also necessarily contain mirror planes parallel to each of
these axes. The explicit proof of this assertion will not be presented here.
All three required properties described above are satisfied by each of
exactly sei;en point groups, namely c%(i), cZ~($), ~ 2 h9( 2g q, e ) ~>3 & $ ) >
D4h(&;$). D ~ h ( $ ; z ) and Oh ($3;). The point group notations used
here are those of Schonflies and the international notations are given in
brackets. Both systems of notation are explained in Appendix A. In sum-
mary. the above results mean that exactly seven different point groups are
possible for spatial lattices. They define the seven crystal systems: tri-
clinic (Ct),monoclinic, ( C 2 h ) , orthorhombic ( D z h ) , trigonal ( D 3 d ) . tetrago-
nal (D*h), hexagonal (DGh), and cubic (Oh).

Bravais lattices

Within a given crystal system, several different types of lattices may exist.
Their common property is that they all have the same point symmetry, but
they may differ otherwise. Figure 1.2 visualizes them by means of their unit
cells. These differences give rise to diflerenf lattice types. The simplest lat-
tices ef a given point symmetry are represented at the far left of each row
in Figure 1.2. They axe called primitive lattices. Even these simple lattices
are not unambiguously &ermined by their point symmetry. If one, for in-
stance, multiplies all lattice vectors by the same real number, i.e. stretches
or compresses the lattices evenly on all sides, the point symmetry remains
unchanged. In less symmetrical crystal system one may even change cer-
tain length relationships or angles between primitive lattice vectors without
disturbing the point symmetry. In the tetragonal system the height of the
10 Chapter 1. Characterization of semiconductors

rectangular parallelepiped in relation to its basis may be changed arbitrar-


ily. Generally speaking, the primitive lattices are determined only up to
continuous point transformations preserving their point symmetries.
Moreover, starting from a primitive lattice one can produce other sets
of regularly ordered points by adding new points to each primitive unit cell
in equivalent positions. As before, equivalent refers to positions which are
either identical or that differ by a lattice vector. If one places the new points
in symmetrical positions, i.e. those which are transformed into themselves
or equivalent points by the point symmetry operations, snch as the centers
of the primitive unit cells, one obtains a new point set having the same
point symmetry as the original lattice. The new point set, in general, no
longer forms a lattice, but may be thought of as the union of several lattices
placed within each other. In some special cases, however, the result may
still be a primitive lattice. Whether this happens or not is a question which
must be explored separately in each case. If the answer is positive, one has
to examine whether the lattice is only another realization of the original
primitive lattice, i.e. whether or not it can be brought back to the original
one by a continuous and symmetry preserving transformation. It turns out
that both cases may occur. If the lattices cannot be transformed into each
other by such a transformation, then this implies that there are two different
types of lattices with the same point symmetry. One calls them different
Bmvais types or Bravais latttices. According to this definition two Bravais
lattices are of the same type if they may be transformed into each other by
a continuous and point symmetry preserving transformation, otherwise they
are Bravais lattices of different types.
As an example, we consider the different Bravais lattices in the case of the
cubic crystal system. If one adds the body centers of the primitive unit-cell-
cubes as additional points to a primitive cubic lattice, the resulting point set
has the cubic point symmetry and it forms again a lattice. The same holds if
the added points are the centers of the faces of the primitive unit-cell-cubes
instead of the body centers. Neither the spacecentered nor the face-centered
cubic lattices can be transformed back to a primitive cubic lattice by means
of a continuous and symmetry preserving transformation, nor can the two
centered lattices be transformed into each other by such a transformation.
Therefore, they represent cubic lattices of two new Bravais types. If one adds
both face and body centers to the primitive cubic lattice, then the new point
set forms again a primitive cubic lattice, however, with a lattice constant
equal to half of that of the original lattice. It may be transformed back to
the original primitive cubic lattice by means of a continuous and symmetry
preserving transformation, thus it does not represent a cubic lattice of a
new Bravais type. Altogether one finds three different cubic Bravais lattices,
the primitive (p), the body-centered (bc), and the face-centered (fc) ones.
Analogous considerations have to be made for the other 6 primitive Iattices.
1.2. Atomic structure of ideal crystals 11

Chic F
Cubic P Cuhic 1

Tetragonst I

pj$,
I
.# .\

@
Triclinic

Monocfinic P Monoclinic 1

Trigonal R Trigonal and hexagonal P

Figure 1.2: Common unit cells of the 14 Bravais lattices.


12 Chapter I , Charactmzation of semiconductors

In this w q ,one finds that, in totul, 14 diEerriit Bravais lattices are possible
in the 7 crystal systems. The assignment ol Iht.sr lattices to the crystal
systems is indicated in Table 1.1. The Bravais lattices themsehps are shown
in Figure 1.2.

1.2.2 Point groups of equivalent directions and crystal classes


The lattice of a crystal serves as a conceptional basis for the illustration of its
translation symmetry. Lattice points do not necessarily have to be occupied
by atoms, which may be bum1 at general points of the primitive unit cells.
Generally, a primitive unit cell contains several atoms which may be either
chemically identical or different. We denote them here by an atom index 1
which t,akes the values 1 - 1 , 2 , . . . , L , where L is the total number of atoms
in a primitive unit cell. For L 2 2 the set of all L atoms is called the basis
uf the crystal. In the case I. = 1 one says t.hat the crystal has no basis.
Ihe position of the 1-th at,om, relative to the corner R of a primitive unit
cell, is drscribed by a vrcbor The position Ri of this atom relative t o the
origin is then given by

Ri=Rti. (1.5)
Without loss of generality one may always set onr of the vectors il, e.g. i l .
equal to zero. If the primitive unit cell canlains only 1 atom, it may be
placed in o m o f the roriiers of the cell. A crystal without basis may thiis be
describd as a laltice whosp points are dl occupied by ahoms. For L 2 2,
the crystal may be generated in such a way that one multiples its crystal
lattice L-fold, then shifts the resulting sublattices relative to t,he first by,
respcctivcly, the vectors 6 , . . . ,?I,, and finally occupies the points of the
shift,ed lattices, respectively, with atoms of t,he species 2,. . . , L.
With only one atom per primitive unit cell! any point symmetry opera-
tion of the lattice will ncccssatily transform the whole crystal into itsdf. For
cryst,alswith basis, however, this is not true in general. For this reason it is
meaningful to consider, besides the point symmetry operations of the crys-
tal lattice, also orthogonal transformations which map physically equivalent
direchns of Ihe crystal into each other, without necessarily bringing the
crystal back onto itself. We explain the meaning of physical equivalence by
using the example of the relation between the vectors of the electric current
density j and the electric field strength E in a crystal. Generally, j is a
non-lincar fimrtioii of E and, because of crystal snisotropy, the direct.ions
of the two vectors may be different. If E and j are transformed from their
original directions reht.ive to the crystal into new ones, without changing
their relative orientation, the relation betwccn the new j and E, in general,
will differ from the relation before rotat,ion. Analogous statements hold for
reflections and rotation-reflections. If tbcre are orthogonal transformations
1.2. Atomic structure of ideal crystals 13

Table 1.1: Symmetry classification of crystals. The following abbreviations are


used p - primit.ive, bc - body centered, fc face centered, bfc - basis face centered.
~

Crystal Holohedral Braiais Crystal Number of space

+
system group lattk class groups

r
tnclinic

monoclinic '2h
c
1

C, ! I
1
1
1

rhombic

trigonal

tetragonal

I
hexagonal

I
)
cubic p bc fc
14 Chapter 1. Characterization of semiconductors

which leave the relation between the two vectors unchanged - which ulti-
mately must be verified experimentally - the original and the transformed
directions are called physically equivalent. The group of orthogonal trans-
formations which map physically equivalent directions of a crystal into each
other, is called the point group of equivalent directions. It defines the crystal
class. Each crystal class corresponds to a particular point group of equivalent
directions. In crystals without basis the point group of equivalent directions
coincides with the holohedral point group. For crystals with a basis this is
no longer true in general, and the point groups of equivalent directions are
generally subgroups of the holohedral groups. There are as many different
point groups of equivalent directions, or crystal classes, as there are different
subgroups of the 7 holohedral groups. Using Appendix A one can easily
show that their number amounts to 32 (see Table 1.1).
Not all crystal classes can be realized in all crystal systems - the point
group which defines the crystal class has to be a subgroup of the holohedral
group of the crystal lattice. Each crystal class is, however, found in at
least one crystal system, several in more than one. In assigning the crystal
classes to the different crystal systems one proceeds as follows: A class which
exists in several systems is attributed to the one with the lowest common
symmetry. In this way one obtains the assignment between crystal systems
and crystal classes shown in Table 1.1.

1.2.3 Space groups and crystal structures


It remains for us to explore the symmetry of the crystal as a whole. This
consists of the set of all rigid displacements which transform the atoms of the
crystal into identical or equivalent positions. It is obvious that symmetry op-
erations which transform equivalent atoms into each other must necessarily
also do the same with physically equivalent directions. The converse, how-
ever, is not always true - after carrying out s rotation, reflection, rotation-
reflection or rotation-inversion which transforms a particular direction into
a physically equivalent one, the atoms of the crystal are not necessarily also
trausformed into equivalent atoms. It may be necessary to add to a rotation
yet another translation parallel to its axis (screw-rotation), or to add to a
reflection a displacement parallel to the mirror plgraetglide-re~ection),or to
add both in the case of rotation-reflections and rotation-inversions. That
one records in this way all conceivable symmetry operations follows from a
theorem proven in Appendix A, which states that every rigid displacement
which is not a pure translation or an orthogonal transformation, must be a
screw-rotation or a glidereflection.
The parallel displacement P; which follows a rotation about an n-fold axis
in a screw-rotation, must be an integer multiple of one n-th of the smallest
lattice vector in the direction of the axis. This follows immediately if one
1.2. Atomic structure of ideal crystals 15

considers the n-th power of the screw-rotation, which is also a symmetry


element as well. This describes an n-fold repetition of the rotation by 27r/n,
i.e. a rotation by 27r, or no rotation at all, followed by the translation n
times +. Since the crystal must pass into an equivalent position by such a
transformation, the vector n times p must be a whole lattice vector. As it
points in the direction of the screw axis, it is necessarily a multiple of the
shortest lattice vector in this direction. In the international notation system
the multiplicity is described by a lower index a f b e d to the symbol for the
rotation axis. This also indicates that the axis is not an ordinary one but a
screw axis (with right-handed thread). The symbol 63, for example, means
a 6-fold screw axis with a parallel displacement by half the shortest lattice
vector in the direction of the axis. For a glidereflection one can similarly
show that the translation in the mirror plane must be a superposition of
multiples of halves of the smallest linearly independent lattice vectors in
this plane.
In summary, we may state that the symmetry operations on a crystal
are of the following types: translations by lattice vectors, proper rotations,
reflections, rotary reflections, rotation-inversions as well as screw-rotations
and glide-reflections. Furthermore, all combinations of these transformations
are allowed. The set of all symmetry operations on a crystal forms a group.
It is called a space group. If the crystal has no screw-rotations or glide-
reflections as symmetry operations, then its space group necessarily contains
the point group of equivalent directions as a subgroup. Such space groups
are called symmorphic. Space groups with screw-rotations or glide reflections
are called non-symmorphic. The latter do not contain the point group of
equivalent directions of the crystal. Each element of a symmorphic space
group is the product of an element of its translation group and an element
of the point group of equivalent directions. The set of all possible space
groups of crystals may be obtained in the following way: One considers,
first of all, crystals of the triclinic system. In this case only the primitive
Bravais lattice is possible. The corresponding space groups may easily be
determined - there are only two. Similarly one proceeds with all other
combinations of crystal classes, Bravais lattices and crystal systems, moving
in the direction of increasing symmetry. At each stage of counting only
the newly occurring space groups are added. In this way one obtains the
numbers indicated in Table 1.1. The total is 230. Each of these 230 possible
space groups corresponds to a particular crystal structure. By specifying its
space group, the structure of a crystal is uniquely determined, except, of
course, for changes of the distances between atoms and of angles between
lines connecting atoms which do not affect the symmetry.
The majority of semiconductors belong to a small selection of the possible
crystal structures or space groups. Five are especially important: Their des-
ignations in the international system are Fd3m (diamond structure), F W m
16 Chapter 1. Characterization of semiconductors

(zincblende st ruct ure) , F m3m (rocksalt st ruct ure) , P63m c (wurt zite st ruc-
ture), and P3121 or P3221 (selenium structure). Each of these crystal struc-
tures is more or less closely tied to a particular material group. We consider
this connection in more detail in the next section, which deals with the
chemical nature of semiconductors.
The five crystal structures above belong to just two different crystal sys-
tems - t h e diamond, zincblende and rocksalt structures belong to the cubic
system and the wurtzite and selenium structures belong to the hexagonal
system. We start with the three cubic structures. Table 1.2 summarizes
their properties

1.2.4 Cubic semiconductor structures


Crystals having diamond, zincblende and rocksalt structures not only belong
to the same crystal system, but also have the same Bravais lattice, namely
the face centered cubic, abbreviated as fcc. The fcc lattice is commonly
described in a Cartesian coordinate system whose unit vectors e,, ey,e, are
taken in the directions of the cubic crystal axes. The lattice constant a is
the distance between the lattice points of a primitive cubic reference lattice,
which is obtained from the fcc lattices by omitting the face centers. The
primitive lattice vectors al,a2,* of the fcc lattice may be chosen in the
form

The lattice constants for a series of semiconductors with fcc lattices are listed
in Table 1.3. We stress that the cubic lattice constant is neither the distance
between two lattice points (it is u / f i for the fcc lattice), nor the distance
between two atoms in a crystal with this structure (which is &/4). For
the volume Ro of the primitive unit cell one obtains the value u3/4 from
equation (1.3). In the case of silicon this yields Ro = 4.00 x cm3. A
silicon crystal of volume 1 cm3, therefore, contains 2.5 x primitive unit
cells.
The lattice does not determine the positions of the atoms. This is done
by the basis of the crystal, in which respect the three cubic semiconductor
structures differ.
1.2. Atomic structure of ideal crystals 17

Table 1.2: Characteristicz of the five most comnion structure types of seniiconduc
tors. 0 < C < h, 0 < n < $. More details are given in the text.

Diamond Zincblende Itocksalt Wiirtzite Seleniuni

F#m P3121
P3221

fy
,3;
4 - 2
43m Gmm 32

Bravais cubic cubic cubic hexagonal hexagonal


lattice fc fc fc P P

Primitive
lattice
vectors

Basis
I

I I
18 Chapter 1. Characterization of semiconductors

Table 1.3: Lattice constants of semiconductors of various types of structures.

1:
Diamond structure

, I
1:
C Si Ge a-Sn
5.43 5.65 6.46
Zincblende structure

AlP RlAs AlSb GaN Gal Cans CaSb Id


4.37 5.47 5.66 6.14 4.54 5.45 5.65 6.13 5.87
InAs InSb ZnS ZnSe ZnTc CdTe HgSe IlgTe
6.05 6.47 5.43 5.66 6.08 6.42 6.08 6.37

Rocksalt structure I I
PbS PbSe PbTe SnTc CdO MgS MgSe
5.93 6.12 6.45 6.3 4.70 5.20 5.46

Wurtzite structure I I
ZnS CdS CdSe MgTe AlN GaN InN
3.82 4.14 4.31 4.53 3.11 3.19 3.54 a
1.64 1.62 1.63 1.62 1.60 1.62 1.61 cla
Selenium structure
Se Te
4.36 4.47 a

1.14 1.33 cl a

Diamond and zincblende structure

For diamond and zincblende structures, the basis consists of two atoms
which, respectively, have the same or different chemical natures. The two
atoms are shifted with respect to each other in the direction of the body
diagonal of the primitive reference cube by a quarter of its side. If the basis
atom 1 is put at a lattice point, then the position of atom 2 is given by
1.2. Atomic structure of ideal crystals 19

Figure 1.3; Spatial model of a crystal having diamond (a) and zincblende structure
(b). In the diamond structure all atoms arc of the same chemical clement, in the
zincblendc structure the atoms are of two different chemical elements.

0 1A (1111
02 8 a IA
3A r, 2 0
0 4B 3A
(J 40
P
6 3 * 5A
60

Cmll cfm

Figure 1.4: Projection of a crystal having diamond and zincblende structure in a


(100) plane (left) and a (111) plane (right). Atoms of the same size and darkness lie
in the same plane. The vertical sequence of atomic layers is indicated on the right
hand side of the projections. The sequence for the (100) plane is repeated after 4
layers, and that for the (111) plane after 6 layers.

Crystals with diamond and zincblende structures may also be described as


a composite of two interpenetrating superposed fcc lattices displaced with
respect to each other by the vector ?2 and with their lattice points occupied
by, respectively, chemically equivalent or different atoms. The geometric
relations are depicted in Figures. 1.3 and 1.4 in, respectively, a 3-dimensional
representation and a plane projection. From Figure 1.4 one can readily
see that for both crystal structures the cubic axes put through an atom
form 4-fold mirror-rotation or inversion rotation axes (which are the same
in this drawing). The body diagonals through an atom are %fold axes of
20 Chapter I . Characterization of semiconductors

Figuw 1.5: Stereograms of the paint groups Oh (left) and 7'6 (right) of, respectively,
the diamond and zinchlmde structiires.

rotation, and the planrs through opposite cube dges are mirror planes. Both
structures are theiefore tramformd into themselves by the point group T d ,
the symmetry group of a tetrahcdron. It follows that their point groups of
equivalent directions are at least T d . For the zincblende structure there are
no further point symmetry operations which transform equivalent directions
into each other. The point group of equivalent directions is therefore T d , and
the space group turns out to be symmorphic. It is denoted by F13m.
In the raw of diamond structure, thew i s yet another symmptry oper-
ation, namely that which transforms the two chemically equivalent atoms
of the basis into cach other. It may be described as reflection in a plane
perpendicular to a cubic axis, say e,, which cuts the connecting line be-
l w r n two atoms at its center, followrd hy a translation in that plane by
+
(e/4)(ex ey). One, therefore, has a glidereflection as additional symmetry
dement. The space group of diamond structure becomw non-symmorphir
in this way The point group of directions of this structure may be obtained
from the point group T d of the zincblende structure by adding the reflections
in planes perpmdicular to the thrw cubic axes. This yields the full cubic
group Oh, as one can easily determine by means of the stereograms of the
Iwo groups rn Figure 1.5 (for a n introduction to the stereograms of point
groups, see Appendix A). The group o h can be generated from t h e tetra-
hedral group in yet anothcr way, namely by adding the inversion operation.
As an element of the space group, inversion must be relative to the site of an
atom and be followed by a tradation through the vector T2 or -?zr depaid-
ing on whether m c considers a 1- or 2-atom. For future reference we bear
in mind that for the diamond and zincblende structures, inversion always
involves an exchange of the two sublattims.
We now extend our discussion to thc neighborhood-relations in the two crys-
21

Figure 1.6: Spatial model of a crystal


with rocksalt structure.

tal structures, the so-called coordmatatm Consider atom 1 of thc Lesis of the
unit cell at R = 0. The basis atom 2 of this cell is one of its nearebt neigh-
bors. If all rotation-reflections are executed with respect to the axis through
atom I, then the basis atom 2 of the primitive unit cell at R = 0 gives
rise t o three more basis atoms of type 2. They are all at the same distance
from atom 1 as thP 2-atom first considered, but naturally they lie in three
other primitive unit cells. This means that each 1-atom has, altogether,
four nearest neighboring 2 atoms. Since it would also have been possible
to start from a 2-atom in this consideration, the mme also holds with re-
spect to a %atom, and, of course, independently of whether t h 2-atom is
of the same chemical nature as the 1-atom (diamond structure) or is not
(zincblende structure). The four nearest neighbor atoms lie at the corners of
a tetrahedron, whose center is occupied by the atom itself (see Figure 1.3).
Their relative positions are given by the vectors (a/4)(1, 1, l),(e/4)(T, 1,l),
(a/4) (1,7,l), (a/4)(1,1,i). Thr distanrc betwrrn nearest neighbor atoms
is &a/4. For silicon this is 2+35 A . The second nearest npighbors belong
to the same sublattice. Therefore, for both structures, they arc atoms of
thP same chemical species. They are located at thc nearest-neighbor lattice
points of the fcc lattice. relative to the central atom, thus their positions are
( u / 2 } ( * e 9 f ef), ( a / 2 ) ( f e y f e z ) ,(a/2)(*e, f ey). This means that
thew are 12 second nearest neighbors in each slructurr.

Rocksalt structure

The basis of this crystal structure consists, like that of zincblende structure,
of two atoms of diffprent chemical nature which arc displaced relative to each
other in the direction of the space diagonal of the reference cube. In contrast
to zincblende structure, the displawmcnl i s not, however, a quarter but half
of the length o the space diagonal of the reference cube, whence
22 Chapter 1. Characterization of semiconductors

(1111

1A 1A
01B 028
4 2A 3A
0 28 48
-
0
5A
0 6B

Figure 1.7: Projection of a crystal having rocksalt structure in the (100) plane
(left) and the (111) plane (right). The vertical sequences of atomic layers is shown
on the right hand side of the projections. In the (LOO) case the stacking repeata
itself after 4 Iayem, and in the (111) case after 6 layers.

Accordingly, a crystal having rocksalt structure can be described as a com-


posite of two fcc lattices, displaced with respect to ~ a c hother by the vector
?'z, and with their lattice points occupied by chemically different atoms.
This structure is illustrated in Figures 1.6 and 1.7. As one can see horn
these drawings, the crystal has the full symmetry of the primitive reference
cube. The point group of equivalent directions is therefore the full cubic
group o h (see Figure 1.5). The space group is symmorphic and is denoted
by FmSm.
In Ihe direction of a cubic axis, the atoms are separated by a distance of
a / 2 , with 1- and 2-atoms alternating. Since shorter distances do not occur,
these atoms are ncarest neighbors. Each atom, therefore, has 6 nearest
neighbors. The second nearest neighbors are the nearest neighboring atoms
of the same fcc sublatticp. Their number amounts to 12, as in the zincblende
structure. The relative positions are also the same as in this structure.

1.2.5 Hexagonal semiconductor structures


The primitive lattice vectors of the hexagonal lattice may be written in the
form

h
ai y- ae,,
a h
a2 - --ex
2
+ -&e,",
a
2
h
ce,.

Here ek, I$, e," are unit vectors of a cubic coordinate system whose z-axis
points in the direction of the c-axis of the hexagonal lattice and whose z-axis
is identifird arbitrarily with one of the three symmetric lattice directions in
the plane perpendicular to the c-axis. The two primitive lattice vectors a1
1.2. Atomic structure of ideal crystals 23

and a2 in this plane have equal lengths and form an angle of 2x13 with
each other. The volume 5-20 of a primitive unit cell is &a2c/2. The lattice
constants a and c of some semiconductors with a hexagonal lattice are listed
in Table 1.3.

Wurtzite structure

The basis of crystals with wurtzite structure consists of 4 atoms. We denote


them by M I , X I , Mz and X 2 . Each pair of these atoms, namely M I , M2,
on the one hand, and XI, X 2 ) on the other hand, are chemically equivalent,
while the two pairs are of different elements M and X. If atom M I is sited
at a lattice point, then the position vectors of the three others are

fx1= cc e," , (1.10)

Here, ( is a parameter which may take any value between 0 and 1/2. A
wurtzite type crystal can be understood as a composite superposition of
four interpenetrating hexagonal lattices, with the lattice points of two of
them occupied by atoms of chemical species M, and two of them by atoms
of species X. The two M-lattices are displaced relative to each other by
F M ~ ,likewise as the two X-lattices. The displacement of the XI-lattice
with respect to the MI-lattice is Fx1. In Figures 1.8 and 1.9 these relations
are illustrated in, respectively, spatial and planar displays. The M-atoms
are shown in black and the X-atoms in white. From these figures one can
easily determine the symmetry of the wurtzite structure. The c-axis through
the center of one of the empty hexagons in Figure 1.9 forms a 6-fold screw
axis 63: a 2n/6-rotation about this axis must be connected with a parallel
displacement by c/2 in order to transform the crystal into itself. This also
mean8 that this axis represents a %fold proper axis of rotation. The axis
contains six inequivalent mirror planes, among them three proper reflection
planes and three glide-reflection planes joined with translations by c/2 in
the direction of the c-axis. The corresponding point group of directions is
c 6 v (see Figure 1.10). The space group of the wurtzite structure is non-
symmorphic and is denoted by P6amc.
Consider the atom M I in the unit cell at R - 0. Its neighbor in
the same unit cell is atom X 1 at the site ?XI. Another neighbor of M I is
the atom Xz at the site -ce,h I Fxz of the unit cell at R - -c$. Since
the c-axis through atom M i is a %fold symmetry axis (sep Figure t.9), one
obtains from Xz two additional X-atoms at the same distance but in other
unit cells. Altogether, there are four neighboring X-atoms. In order that
all four be at the same distance from M I , and, correspondingly, all four be
24 Chapter 1 . Characterization of semiconductors

Figure 1.8: Spatial model of a crystal


with ideal wurtzite structure.

0 1 A
028
3A
0 4B

Figure 1.9; Projectionof a crystal with Figure 1.10: Stereograms of the


ideal wurtzite structure in the plane point group c6~of the wurtzite
normal to the c-axis. The vertical se- structure.
quence of atoniic layers is shown on the
right herid side.

nearest neighbors of M I , the distance 1 7x1 I between M I and X I must be


equal to the distance I -ce," t F x 2 I between Mi and X2 in the unit cell at
R = -ce!. This results in the condition

c = , 1t 9 ( ;1) . (1.11)

If it is fiilfilled, the atom Mi has four nearest neighboru of chemical species


X . The same statement is true for M I , with the interchange of M and
X , dito for X I and Xa. In semiconductor crystals of the wurtzite type,
the condition (1.11) is always satisfied almosl perfectly. This means that
the four nearest neighbors of an atom are sited at the corners of a slightly
deformed tetrahedron which surrounds the central atom symmetrically, and
1.2. Atomic structure of ideal crystals 25

which is compressed or stretched in the direction of the c-axis. The degree


of deformation depends on the ratio cla. For the deformation to vanish, the
distance between the M I - and XI-lattice planes perpendicular to the c-axis
must be three times the distance between the X I - and M2-lattice planes
perpendicular to this axis, just as in the zincblende structure. This yields
< <
= 3(1/2-<) or = 318 which, taken together with equation (l.ll),results
in

(z) fi=
= 1.633. (1.12)

Table 1.2 shows that all wurtzite type semiconductors listed there also satisfy
this condition quite well. This indicates that the ordering of the nearest
neighbors in the wurtzite structure is almost the same as in the zincblende
structure. If a wurtzite crystal exactly obeys conditions (1.11) and (1.12),
one says that it has an ideal w u r t z i t e structure. The distance between two
nearest neighbor atoms in the zincblende structure is &all4 with a the
cubic lattice constant. The corresponding distance in the wurtzite structure
is Cc = (3/8)@ a. For the two distances to be equal, the equation

a= h a (1.13)

must hold. This result means that in a crystal having wurtzite structure
which fuEUs conditions (1.11) and (1.12), the nearest neighbor atoms are
positioned at exactly the same sites as in a crystal having zincblende struc-
ture with lattice constant 4 a. If one multiplies the lattice constants a,
of thc wurtzitc type crystals in Table 1.2 by 4,then one in fact obtains
values which fit well witahthe cubic lat,t,iccconstants of the zincblendc struc-
ture crystals of this table. The volume &a2c/2 of a primitive unit cell oi
the ideal wurtzitc structure is d 3 / 2 , which is twice that of thc zincblende
reference structure given by 8 1 4 .
Of course, this does not at all mean that the wurtzite structurc may
be traced back to lhe zincblende slruclure, Indeed, the actual positions of
the second nearest neighbors differ in the two crystal structures, although
they are all at the same distances a = a/& This may be seen most easily
from the projections of the two crystals on the plane perpendicular to the
c-axis (see Figure 1.4) or to the space diagonals e, -t e$ - t e, (see Figure
1.9). In the wurtzite structure the vertical stacking repeats itself after two
double layers of nearest neighbor atoms, but in the zincblende structure only
after three. The wurtzik structure can be set up, therefore, by stacking one
upon another two different double layers A and B, the zincblende structiire
by stacking three double layers A, B,C (see Figure 1.11). The uppermost
double layer A of the zincbbnde structure in Figure 1.llb coincides with the
26 Chapter 1. Characterizrrtion of semiconductors

A B
-
* A - A
-
-B - B
-A - C
-
. B -A - B
-C
:

Figure 1.11: Construction of a crystal having wurtzite structure (top) and


zincblende structure (bottom) by stacking of double layers of nearest-neighbor
atoms. Atoms shown by full circles form the upper part of a double layer, and
atoms shown by open circles form the lower part of such a layer. In stacking, the
nets shown in each of the double layers must be placed at the same positions. The
stacking sequence is depicted on the upper right. M h e r explanations are given in
the text.

uppermost double layer A of the wurtzite structure in Figure L l l a , while


B and C differ from B. By examining the layers in Figure 1.11 we also
note the different locations of some of the second nearest neighbor atoms of
the two crystal structures. Among the 12 second nearest neighbors of an
atom in an A = A-layer, in the ideal wurtzite structure, three are found in
the upper B-layer and three in the lower B-layer, while in the zincblende
structure three second nearest neighbors are in the C-layer above and three
in the B-layer below. In Figure 1.11 this fact is illustrated by emphasis
on the second nearest neighbor atoms mentioned. Those in double layer B
are not in the same positions as those in double layer B.Thus the second
nearest neighbor sites are partially different in the wurtzite and the ideal
zincblende structures.
A comparison of the two crystal structures shows that the same ordering
of the nearest neighbor atoms may be consistent with different orderings of
more remote atoms. From this observation we may conclude that, in a solid,
1.2. Atomic structure of ideal crystals 27

Figure 1.12: Spatial model of a crystal


with selenium structure.

the long-range order could be perturbed without significantly changing the


short-range order of the corresponding crystal. This is just what is observed
in amorphous silicon and other amorphous semiconductors.

Selenium structure

The basis of trigonal selenium and tellurium crystals consists of three atoms.
The vectors of the basis may be written in the form

2c
~3 = -2craez +-
3 e,h, (1.14)

where a is a parameter ranging between 0 and 112. The atoms are ordered
on parallel spirals, in which successive atomic positions are rotated with
respect to each other by an angle of 120' (see Figure 1.12). The space
group can be determined from the projection of the crystal on the plane
perpendicular to the c-axis (see Figure 1.13). It contains one %fold screw
axis parallel to the c-axis which goes through a center of one of the empty
triangles in Figure 1.13. A rotation by 120 or 240' must be followed by
a displacement by c/3 or 2c/3 in the direction of the c-axis, to transform
the crystal into itself. Beside these screw operations, rotations are allowed
with respect to the three inequivalent 2-fold axes perpendicular to the c-
axis. These rotation axes pass through the %fold screw axis and through
one of the three surrounding atoms. The symmetry operations mentioned
correspond to the space group P312. Beside this, the selenium structure
may also have symmetry corresponding to the space group P322, where the
spiral differing from P312 has not a right-handed thread, but a left-handed
one. The point group of directions is, in both cases, D3 (see Figure 1.14).
Concluding this discussion of the characterization of the most common
semiconductor crystal structures, it should be emphasized that these struc-
28 Chapter 1 . Characterization of semiconductors

(00011

o c c 3

Figure 1.13: Projection of a ciystal Figure 1.14: Stereograms of the


having selenium structure on the plane point group D3 of the selenium
normal to the c-axis. The vertical se- structure.
quence of atomic layers is shown on the
right hand side.

tures apply to the ideal case of infinite, structurally perfect, and chemically
absolutely pure materials. Real crystals deviate from this ideal case in vary-
ing degrees. There are structural defects such as stacking faults, step- or
screw-dislocations, vacancies, or atoms on interstitial crystal sites. Such de-
fects will be considered in detail in Chapter 3, in the context of the electronic
structure of yprturbrrl semiconductors. In the next section we will discuss
the chemical composition of idea1 semiconductor crystals.

1.3 Chemical nature of semiconductors. Material


classes
According to their chemical nature, most semiconductors are inorganic ma-
terials. Examples of organic semiconductors indude anthracene, naphthaline
and polyacetylcne. In thiv book, .we restrict our considerations to inorganic
semiconductors, because these are much bettcr understood and haw much
greater twhnolagical iniporttznce than semiconductors of the organic type.
Nevertheless, organic semiconductors are also of interest for science and tcch-
nology.
The best way to get an overview of the different &sues of inorganic
semiconducting materials is to examine thP periodic table of elements. In
'l'able 1.4. a part of the periodic table is shown with which many elemental
and compound semiconductors are associated. W e start with the clernenld
semiconductors of group IV.
1,3. Chemical nature of semiconductors. Material classes 29

3 c N 0
Mg A1 Si P S
Zn Ga Ge As Se
Cd In Sn Sb 're
Hg T1 Pb Bi

Table 1.4: Part of the periodic table t o which Inany element.al and compound
semiconductors are associated.

1.3.1 Group IV elemental semiconductors

Semiconductors of this group are, starting at the top of column IV? carbon
(C) in the form as diamond, silicon (Si), germanium (Ge), and tin (Sn) in
its gray modification known as a-Sn. All cited semiconductors of this group
crystallize into the structure of diamond. They differ from each other in
regard to their position Let,weezl metals and insulators. Diamond behaves
much like an insulator! and tin, on the contrary, much like a metal. Retween
them one has the two typical semicondnctors silicon and germanium. Both
mat.&& play an essential role in micro- and optoelectronics. It is weU-
known that silicon is the material of preference in microelectronics and,
with this, one of the most important materials in dl modern technology.
Beside the pure elemental semiconductors Si a i d Ge, alloys of bath ma-
terials also have semiconducting properties. The notation for such alloys is
(Si,Ge), or Sil-,Ge, to indicate the rpspcctive mole fractions 1 - z and z
of the two alloy components. According to their structure, the alloys may
be identified as mixed crystals. These are crystals with the same geomcb-
ric order of atom sites as in the case of the alloy components here that
~

of diamond in both CBSW. The regular crystal sites are occupied, however.
randamly by Si or Ge atoms. T'he probability of finding a particular el*
ment on a given site is determined hy the mole frclctiou of this element. in
t,he alloy. In the case of Sil-,Ce, the probability of finding Si is 1 - 2.
and of finding Ge is x. The lattice constants of inixcd cryst.als, in many
30 Chapter 1 . Characterization of semiconductors

cases, follow from the lattice constants of the constituents by means of En-
ear interpolation. This result is known as Vegards rule. The cubic lattice
constant a(si,ce)of a Sil-,Ge, alloy, for example, is given by the relation
~ ( s i= +
, ~(1)- z)asi Z U G ~ ,where asi and a& are the cubic lattice con-
stants of Si and Ge, respectively. A mixed crystal strongly deviates from
an ideal crystal, but the semiconducting property of the two pure crystals
is often retained in the mixed crystal. This is important, and by no means
obvious. It was recognized at the beginning of the sixties in (Si, Ge) alloys.
Later it was also shown to be true in many other material combinations.
The discovery of semiconducting alloys significantly expanded the family of
semiconductor materials. It became possible to tailor materials with desired
properties, just by varying the alloy elements and mole fractions.
The common feature of the elements of the main group IV of the peri-
odic table is that there are four electrons in the outer shell of their electron
clouds, the so-called valence shell. The primitive unit cell, which here has
two atoms, therefore contains 8 electrons. Later we will prove that semi-
conducting properties are related to the crystal structure and the number of
valence electrons per primitive unit cell. Using this result, one may conclude
that compounds of elements from main groups III and V, and from main
groups 11 and VI of the periodic table should also be semiconductors. pro-
vided they have zincblende structure, since the crystals of these compounds
also have 8 electrons per primitive unit cell. Actually the GI-V compounds
(with a few exceptions) do crystallize into the zincblende structure, In the
case of ILVI-compounds the wurtzite structure can also occur, besides the
zincblende structure. The two structures are, however, very similar, as has
been pointed out above, so that semiconducting behavior may be expected
also in their case.

1.3.2 111-V semiconductors


The conjecture that the 111-V compounds should be semiconductors is amply
confirmed. Following the group III and group V columns of the periodic table
from the top down, one obtains the following compound semiconductors:
BN, BP, BAS, AlK, Alp, AlAs, AlSb, GaN, Gap, GaAs, GaSb, ZnN, InP,
InAs and InSb. Except for the nitrides all these compounds crystallize into
the zincblende structure. The nitrides are stable in the wurtzite structure,
BY and GaN also have metastable zincblende phases. Just, as in the case of
the elemental semiconductors of group IV, the mixed crystals made of binary
111-V compounds also have semiconducting properties. Examples are (Ga,
Al)As, Ga(As, P), (In,Ga)As and (In,Ga)(As,P). The principal applications
of the 111-V semiconductors and their alloys lie in the field of optoelectronics.
(Ga,Al)As and Ga(As,P) are used, or example, in light emitting diodes
(LEDs) and laser diodes for the near infrared t o green spectral region. GaN
1.3. Chemical nature of semiconductors. Material classes 31

is a promising material for blue LEDs and laser diodes. The quaternary alloy
system (Ga, In)(As, P ) is used in making laserx and pholodiodes for optical
fiber communications at 1.55 p m wavelengths, which provides maximum
transmission for SiOa-based fibers. GaAs and, lately, also (Gtr,Al)As are
employed for transistors with extremely short time delay. BN and GaN can
be used for electronic devices to be operated at high temperatures.

1.3.3 11-VI semiconductors

This group includes compounds like ZnS, ZnSe, ZnTe, CdTe, HgSe and HgTe,
which crystallize into the zincblende structure, and compounds like CdS,
CdSe and MgTe which have the wurtzite structure. For most of these mate-
rials the one of the two structures which has not been mentioned corresponds
to a metastable phase, and for ZnS both structures are, in fact, stable. The
cubic phase of ZnS is the mineral zincblende, and the hexagonal phase is
the mineral ururtzite. It i s from these mineral names that the designations
of the two crystal structures are derived. CdO, MgS and MgSe form crys-
tals with rocksalt structure, HgS (vermilion) has a rocksalt-like structure.
As in the case of the 111-V compounds a large number of semiconducting
alloys may also be realized from the IT-V1 compounds, e.g. (Hg,Cd)Te,
Zn(S,Se), Cd(S,Se) and others. Up to now, technical applications involve
mainly (Hg,Cd)Te. It is utilized for detectors of radiation in the medium
infrared spectral region, which is of particular interest since the emission
maximum of the thermal radiation of bodies lies in this region for a broad
range of temperatures. Some semiconductors of this material class, e.g. CdS,
have very good photoelectric properties and are used accordingly. Other such
compounds, like ZnS, are good luminescent materials. They make it pos-
sible to fabricate largearea electroluminescence displays, an alternative to
the traditional vacuum-tube. The operation of blue-green LEDs and injec-
tion lasers made of 11-VI compound semiconductors was demonstrated very
recently.

1.3.4 Group VI elemental semiconductors

Selenium and tellurium are likewise good semiconductors. Both crystallize


into the same trigonal chain structure, which is characteristic of these mate-
rials (see Table 1.2). Selenium has a long history of practical use. Before the
introduction of power rectifiers based on silicon, selenium ones were used.
Today, above all, the photoelectric properties of selenium are exploited. This
applies, for example, to photocopying where the photosensitive layer of the
photostatic cylinder may be made of selenium or a selenium alloy.
32 Chapter 1. Characterization of semiconductors

1.3.5 IV-VI semiconductors


EIements of groups IV and VI can also form semiconducting compounds.
Examples are PbS, PbSe, PbTe and SnTe, which all crystallize into the
rocksalt structure. Important semiconducting alloys of such compounds are
(Pb,Sn)Te and Pb(S,Se). These semiconductors are now used primarily for
the generation and detection of radiation in the medium infrared spectral
region. In addition, there is a second group of semiconducting IV-IrI com-
pounds which display an orthorhombic structure, including GeS, GeSe, SnS
and SnSe. GeTe is an exception as it crystallizes into a rhombohedra1 struc-
ture.

1.3.6 Other compound semiconductors


It is to be expected that a combination of two elements of group IV will be a
semiconductor. It seems that S i c is the only stable compound of this kind.
It is one of the longest known semiconductors of all. Electroluminescence
was observed in this material as earlier as 1907. S i c OCCUIS in many crystal
structures, among them the zincblende and wurtzite structures. After a long
period of only moderate interest in Sic, much activity is now being devoted
to this materiai. The reason is the good thermal stability of Sic. which
makes it possible to fabricate electronic devices for high temperature and
high power applications.
The elements of groups I1 and V may form semiconducting 1 I 3 - v ~com-
pounds. Examples are ZmAsz and CdsAsz, which possess both tetragonal
and cubic crystalline phases.
One also finds semiconducting properties in compounds composed of el-
ements of groups V and VI with the stoichiometric composition V2-VI3. In
particular, the oxides, sulfides. and selenides of the semimetals As, Sb, and
Bi are among them. With regard to crystal structure, a relatively large
diversity exists among these compounds. Different structure types of the
trigonal, orthorhombic and monoclinic crystal systems have been observed.
The very good thermoelectric properties of Bi2Se3 and BizTq have long
been used for cooling elements.
Finally, one also finds semiconductors among III-VI compounds. Exam-
ples are G ~ q S e 3GazTe3
~ or In2Se3, which display a 2:3-stoichiometry. They
crystallize into so-called defect stmctuws. These are modifications of cer-
tain basic crystallographic structures as, for example, of the zincblende or
wurtzite type, which distinguish themselves by having a (ordered or disor-
dered) network of cation vacancies. These vacancies are necessary to make
possible a stoichiometry which deviates h m that of the basic structure.
Among the semiconducting 111-VI compounds, however, there are also those
with 1:l-stoichiometry. The layer-structures GaS, GaSe, and GaTe are ex-
1.4. Macroscopic properties and their microscopic implications 33

amples. A semiconductor of this group displaying ferromagnetic properties


is EuSe.
So far we have exclusively considered bznary compound semiconductors.
Ilowever, thcrc are also semiconductors generated from ternary and qua
ternary compounds. Among the ternary I-IIEVI compounds, one has, in
particular, the I-111-VTz materials which crystallize, such as CuGaSz, into
the chalcopyrite structure belonging to the hexagonal crystal system. Semi-
conductors are also counted among the II-IV-V2 compounds, ZnSiPa as an
example. The crystal structure is also of chalcopyrite type.
Concluding this survey of the most important semiconducting material
classes, we add two more general remarks. Firstly, it has to b r remeinherd
that we have stated the chemical composition of ideal semiconductor crys-
tals, meaning, in particular, chemically pure materials. Absolute chemical
purity, however, occiirs in nature just as seldom as absolute crystallographic
perfection. Each real semiconductor material contains chemical impurities.
In order that crystals composed of such materials really be semiconductors,
their impurity concentrations must not to be too large - bear in mind what
has been said about this point in the introduction. On the other hand,
semiconductors can not to be too pure. In order to have a sufficiently large
electric conductivity, they must often be intentionally polluted with certain
chemical elements. We will undertake this discussion in greater detail in the
next subsection.
Secondly, it is striking that the semiconducting substance classes which
were defined pritnarily by their chemical nature also display, as a riile, the
same crystallographic structure. The reason for this relation between chem-
ical composition and atomic structure is by no means obvious. It is closely
connected with the electronic structure of crystals which we will treat in
Chapter 2.

1.4 Macroscopic prapert ies and their microscopic


irnplicat ions
In this section we describe characteristic macroscopic properties of semi-
conductors and w e them as a guide to reveal some important microscopic
features of these materials as, for example, the distribution of the allowed
electron energy levels in the form of bands being separated by forbidden en-
ergy regions. Although the microscopic model of a semiconductor obtained
in this way cannot be more than a lucky guess, it will stimulate our intu-
ition later in this book when we are going to derive the microscopic physics
of semiconductors, including their band structure, from first principles.
34 Chapter 1. Characterization of semiconductors

Table 1.5: Electrical conductivity 0 and charge carrier concentrations for metals,
serniconduct.ors,and insulators.

Material class Metal Semiconductor Insulator

lo3.. < 10-10


n jcm-3 1021 . . . 1o1O

1.4.1 Electrical conductivity


As already mentioned, the term semiconductor refers t o a characteristic
macroscopic property of these materials, namely their electrical conductiv-
ity, The latter is defined in terms of the proportionality factor between the
current density j and the electric field E in Ohms law

j=uE (1.15)
Here, it is assumed that anisotropies either do not exist or are negligible.
The following example gives an idea of the order of magnitude of the con-
ductivity: Consider a cube with edge length of 1 cm, and apply a voltage of
1 V between two of its opposite planes. Then a current of 1 A will flow if
the cube material has a conductivity of (T = 1 ern-'. The order of mag-
nitude ranges spanned by the U-values of the three material classes metals,
semiconductors, and insulators are assembled in Table 1.5. The striking
features of this table are the large changes both between the three classes of
substances as well as within the semiconductor material class. The question
arises how these sharp changes may be understood 011a physical basis.
To construct an explanation, we use the known representation of the
electrical conductivity IT as the product of the electron concentration n, the
mobility ,u, and the electric charge -e of an electron,

u = enp. (1.16)
The mobility p is defined as the ratio of the magnitude of the average ve-
locit,y of an electron to the strength of the electric field driving it (for a
derivation see Chapter 5). Formula (1.16) for the Conductivity o shows that
its variations may, in principle, be caused either by different values of the
mobility p and/or by different electron concentrations n.
The mobility is actually determined by the scattering processes which
electrons undergo due to the perturbations of the crystal lattice. One ex-
1.4. IlIacrascopic properties and their microscopic implications 35

pecks, of course, differences with respect to the strengths of these processes


in metals, semiconductors, and insulators, but variations by several orders of
magnitude are very unlikely. The orders of magnitude of the mobilities ME,
psc and in normal metals, semiconductors and insulators, respectively,
should be more or less the same:

P M E E? P S C = P I S . (1.17)
This conjecture is confirmed by experiment, and consequently it follows that
the large conductivity differences between the three material groups are es-
sentially due to their different electron concentrations. If one assumes for
metals one free electron per atom, which amounts to approximately 10
electrons per m 3 , then the conductivity values in Table 1.5 yield the elec-
tron concentrations of semiconductors and insulators listed in this table.

1.4.2 Dependence of conductivity on the semiconductor state


The large range of conductivity values of semiconductors, between lo-
0-l c m - l and lo3 R- m- , does not arise, as one might suppose, from the
chemical diversity of semiconductors, but is primarily due to the fact that a
given semiconductor material covers the entire range of cr-values by itself if
its macroscopic state is varied. By state we mean, firstly, the temperature,
and secondly the Concentrations of certain impurities.
Considering first the change of the conductivity with temperature, we
note that it is particularly pronounced if the semiconductor contains prac-
tically no impurities - one c d s this an int~ninsicsemiconductor. In Figure
1.15 the dependence of the conductivity on temperature is shown for very
pure silicon and germanium samples. The conductivity increases with rising
temperature according to the exponential law

(1.18)
Here, rrg is a factor which depends only weakly on temperature, k is the
Boltzmann constant and Eg is an energy of about 0.7 eV in the case of ger-
manium and 1.1el in the case of silicon. Formula (1.18) also describes the
temperature dependence of the conductivity for other pure semiconductor
materials fairly well at sufficiently high temperatures, save that the values
for cro and E, must be changed. In Table 1.6 the &.-values are listed for
some import ant semiconductors.
There is a second way of varying the conductivity of a semiconductor
over a wide range, namely to intentionally pollute it with atoms of certain
chemical elements. One calls this process doping wzth impup.ity atoms. In
this context, the doped semiconductor materials are called extrinsic semi-
conductors. As an example we choose silicon, doped with arsenic a t o m
36 Chapter 1. Characterization of semiconductors

Figure 1.15: Temperature depen- Figure 1.16: Dependence of the conduc


dence of the conductivity a of very tivit.y u of As-doped Si a t 1' = 300 X
pure (intrinsic) Ge and Si. (Ajter an tlic arsenic concentration ~VD. (After
Epzianov, 1979.) Morin and Makta, 1954.1

of concentration N D . In Figure 1.16 the change of conductivity is shown


with N o varying from 1014 up to crnA3. At temperature 300 K cor-
raponding to the data of Figure 1.16, the conductivity rises monotonically
with doping ronrrntration. Again, a change of several orders of magnitude
i s exhibited.

Table 1.6: E,-values for several semiconductor materials at 300 K (in eV).(Aftar
Lundoldt-Bumstein, 1982.)

Material C: Si Ge ,O-SiC BN AlN AIP AlAs AlSb


E, 1 5.5 1.11 0.66 2.2 6.2 6.28 2.45 3.14 1.63

Material GaN GaP GaAs (Mh InN IriP Iiub IrlSb ZnS

E, 3.30 2.27 1.43 0.71 1.95 1.26 0.36 0.18 3.56

Material ZnSe CdS CdTe HgTe PbS PbSe PbTe Se Te

E, 2.67 2.50 1.43 0.00 0.37 0.26 0.30 1.8 0.33


1.4. Macroscopic properties and their microscopic implications 37

Figure 1.17: Temperature de-


pendence of conductivity u of
As-doped silicon for different ar-
senic concentrations N D : 1 -
1.75 x 1014, z - 2.1 1015,3-
1.75 x 101*,4-1.3 x 1017,5-2.2 x
101B.6-2.7x lo1 cm . (After
Morin and Maita, 195d.}

The trmperalurr dependence of the conductivity is substantially weaker for


doped silicon than for pure as may be seen from Figure 1.17. Each curv-e
in this figure corresponds to a particular concentration of arsenic atoms.
The temperature rise becomes weaker with increasing doping, and for the
highest doping it completely vanishes. At higher temperatures, it even turns
downward in a decrease that is more pronounced for purer materials. The
rise of curve 1 above 400 K indicates thcrt the corresponding [relatively pure)
material starts to behave like an intrinsic semiconductor (see Figure 1.15).

The origin of the difference in temperature characteristics of the con-


ductivity of pure and d o p d semiconductors remains to be addressed. In
this regard, it will be helpful to discuss yet another rnacrosropic property of
semiconductors, namely optical absorption.
38 Chapter 1 , Characterization of semiconductors

Figure 1.18: Absorption coefficient


a of Si BS II fimction of photon energy f
Bw. (Ajter Philipp and Taft, 1960.)

100 -'

' ' '


10-2~
1 2
I
4
I
102040
hw/eV
m m
-
I

1.4.3 Optical absorption spectrum and the band model of


semiconductors

The measured values of the absorption coefficient of silicon are displayed in


Figure 1.18 showing a ( h w ) as a function of the photon energy hw. The par-
ticular shape of this spectrum, namely the vanishing of a below a threshold
energy, the steep rise at this energy up to valurb which are larger by orders
of magnitude, and the weak decrease from these high values with further in-
crease of photon energy, i s characteristic of all semiconductors. In simplified
terms, one may say that semiconductors are transparent below the threshold
photon energy, and absorbing above it. The values of the optical threshold
energies coincide so well with the Eg-values known from the temperature
dependence of the electrical conductivity (see Table L6), that an intimate
relation i s strongly suggested. Below we will analyze this relation in greater
detail.
Consider a semiconductor in thermodynamic equilibrium at absolute zero
temperature, and let E, be the maximum energy which an electron may have
m such a semiconductor. If an electron of that energy absorbs a photon of
+
energy h w , lhen its energy increases to the value E , hw. The absorption
can, however, only occur if the energy E, +
hw is in the allowed range
of electron energies. It is well-known that in quantum mechanics not all
energies arc allowed, but only specific energies which are eigenvalues of the
Hamiltonian. For atoms these are the energies corresponding to the various
Bohr orbitals. Thus, in semiconductors, absorption may be non-vanishing
abovc h'g only if the electrons can take energy values E > E , - E* E,, +
i.e. if the energy values E' > E, are allowed. In turn, the vaniahing of
absorption for E < E , can be explained if the energy values between E,
and & are forbidden. Such a distribution of allow4 energy values is shown
schematically in Figure 1.19. The representation of these values as a function
I .4. Macroscopic properties and their microscopicimpIications 39

Valence bond

O0 X--- HL
0 1
f [El--)

Figure 1.19: Ordering of the allowed Figure 1.20: Fermi distribution


e n e r a v d i m of an ideal semicnnduc- function f ( E ) as a function of energy
tor (band model). T h e spatial ex- at different temperat#ur&.The energy
tension of the corresponding electron unit is E,, and EF has been set at
states is also illtlstrnted. E,P.

of electron coordinate is intended to illustrate the degree of localization of


the electron. A straight line across the whole crystal from 0 to d means that
the electron at this energy is spread out over the whole crystal. The regions
with closely clustered allowed energy values and electron states extended
over the whole crystal are called energy bands. The lower band in Figure
1.19 is the valcrice band, and the upper is the conduction band. Between
these two bands lies a region of forbidden energy values, called the forbidden
zone or the energy gap.
The existence of energy bands done is not in itself siifficient, however,
for the explanation of the observed absorption spectrum of semiconductors.
For this it is also necessary that the valence band be occupied by electrons
and the conduction band be empty. If there were no electrons available with
energies in the valence band, then there would &o be no electrons to absorb
a photon to make the transition to the conduction band possible. If, on the
other hand, all states of the conduction band were to be occupied, then no
electron could be excited tu this band, and no photon absorb4 in such a
transition since, according t o the P a d exclusion principle, the conduction
band could host no further electrons.
40 Chapter I , Characterization of semiconductors

0 L 0 L
X- X 4

Figure 1.21: Distribution of electrons of an ideal semiconductor over the valence


and conduction bands. On the left for T = 0 K , on the right for T > 0 K.

The population of energy levels by electrons at equilibrium is determined


by the lkrrnz distribution f i n r t i o n f ( B ) . This function represents the prob-
ability of finding an electron in an energy levcl E , and it is given by the
expression

(1.19)

Here E F is the so-calld Fwmz mevyy. Like temperature, the Fermi en-
ergy is an intensive thermodynamic state variable, namely the free enthalpy
per particle or the chemical potential. A more detailed treatment of the
Fermi distribution function will he given in Chapter 4. The probability of
orcupation of a particular energy value E: depends decisively on the relative
position of E with respect to the Fermi energy. If E < E F holds, and in
addition J E- E F J>> kT,then f ( R ) essentially has the value 1. If, on the
contrary, E > E F and again IE - EFI >> k T , then f ( E ) is approximately
zero. The shape of f(E)is shown schematically in Figure 1.20. The width
of the e n e r a region where the transition horn 1 to 0 takes place is cf the
order of magnitude kT. The lower the temperature, the more abrupt this
transition becomes. At T = 0 K thc transition is steplike. Qccesionally, one
says that f(E)has the form of an iceblock at T 0 K , which melts at
higher temperatures. 111 Figure 1.20 we have also implied that Pi, lies below
E F and E, above E F , in order for the valence bend to be almost completely
occupied and the conduction band to be almost empty. This means that
1.4. hfacroscopic properties and their microscopic implications 41

Figure 1.22: Absorption of photons by


electron transitions from vaIence to con-
duction band.

0 L
X--L

the Fermi energy must be located within the energy gap between Ev and
E,, sufficiently far away from the two band edges (measured in units of kT).
In this. we have described an essential microscopic property of pure semi-
conductors. Their Fermi level lies in the energy gap between the valence
and conduction bands. The electrons are distributed over the two bands as
shown in Figure 1.21. This knowledge provides a complete explanation of
the absorption spectrum in Figure 1.18. The explanation is illustrated in
Figure 1.22 and needs no further comment.
The differences in the properties of semiconductors and metals may be
traced back, in essence, to the different positions of the Fermi levels in these
two different types of materials. In metals the Fermi level lies within the
conduction band. Insulators do not differ from semiconductors qualitatively
with respect to their Fermi level positions, i.e. the Fermi levels are found
in the energy gap in both cases, but the gap of insulators is typically larger
than that of semiconductors. If Eg > 3.5 e V , as a rule, one has an insulator.

1.4.4 Electrical conductivity in the band model

We will now show that the temperature dependence of the electric con-
ductivity of pme semiconductors also follows from the energy band scheme
discussed above and the position of the Fermi level in it. Again, we employ
the relation u = enp. The mobility p can be assumed to be a relatively
weakly varying function of temperature, T . This means that the strong ex-
ponential T-dependence of CT must be due to the electron concentration IZ,
42 Chapter 1. Characterization of semiconductors

and comparison with equation (1.18j yields, approximately

(1.20)

with n o as a factor of dimension ~ r n -depending


~ weakly on temperature.
As we know, the electrons of intrinsic semiconductors are statistically dip
tributed over the valence and conduction bands in accordance with the Fermi
distribution function f(E). At T 0 K the valence band is fully occupied,
7

and the conduction band is completely empty. The electrons in the fully oc-
cupied valence band cannot contribute to carrier transport. The reason for
this is again the Pauli principlc, according t o which an electron state may be
occupied by only one electron. Since all valence states are occupied, no stste
change is possible in the valence band by redistribut,ing the electrons. This
means that the T = 0 K state will remain unchanged in 8x1 electric Eeld, and
consequently no current will flow. However, a current will arise from the rel-
atively few clectrons which, according to thc Fermi distribution function: are
populating the conduction band at ftnite temperature. Their concentration
can be easily calculated. For a pure semiconductor Ec - EF >> 6T holds.
Thus, in the Fermi distribution function j ( E ) of ecpation (l.l!J),out3 m a y
neglect the 1 of the denominator for emrgies E in the conduction band,
i.e. energies with E > EC Iu so doing, one obtains, approximately, the
Boltzmann distribution function

(1.21)

The concentration of energy levels in the conduction band having energies


+
between E and E dE is p c ( E ) d E , where pc(E) is the so-called density
of states of the conduction band, which describes the number of states per
unit energy and unit volume. In Chapter 2 we will calculate this quantity
explicitly, here its mere existence suffices. In terms of p , ( E ) one may write
the electron concentration n of the conduction band in the form

(1.22)

If one substitutes f(E)from (1.21) into the integrand of (1.221, it follows


that

(1.23)
where

(1.24)
1.4. Macroscopic properties and their microscopic implications 43

may be understood as the effectiae density of states of the conduction band.


To ensure the consistency of the expressions (1.20) and (1.23) for n, we
identify no = N , and ( E c - E F )= Eg/2,or

1
E F = Ev + -E
2 g '
(1.25)

i.e. the Ferini level lies exactly in the middle between the valence and con-
diistion bands. Strictly speaking, this is only true for T = 0 K , as we will
see later. The corrections at higher temperatures are, however, quite smaI1,
and they can bc ignored here. Another inaccuracy of the above considera-
tion that. we have ignored is that, at, finite temperatures, the valence band
population also changes. so that no longer are all states of this band ocm-
pied. Under these circumstances t.he electrons of the valence band also make
a contribution to the current. The temperature dependence of this contri-
bution is roughly the same as that of the conduction band. The electric
charge transport due to the electrons of a not-completely-occupied valence
band will be considered more fully later. It is connetted with a remarkable
observation of the Hall effect, which we will discuss below.
However, we will first clarify how one may understand the strong change
of conductivity with impurity concentration in extrinsic semiconductors. For
this purpose we again m e the energy band model of Figure 1.19 for an ides1
semiconductor. If arsenic impurity atoms are present in a silicon crystal,
then this model has to be altered. Wc will later prove (see Chapter 3)
that each of the arsenic impurity atoms g i v ~ rise to an energy level in the
forbidden gap, just below the conduct.ian hand edge, and that the electrons
in these levels are localized at the site of the impurity atom. For this reason,
we have marked these levels in Figure 1.23 by short line segments.
At temperature T = 0 K each of these levels is occupied by one dec-
tron, namely the fifth valence elcctron of an arsenic atom replacing a silicon
atom having only four valence electrons (see Figure 1.23). If temperature
i s increased, these electrons are excited into the conduction band. In this
way, bound electrons which formerly could not participate in carrier trans-
port, become freely mobile electrons (see Figure 1.24) which can contribute
to transport. If the concentration N D of arsenic atoms is not too large: and
the temperature 7' not too law, then practically all arsenic atoms are ionized
and the concentration n of free electrons is equal to the that of the arsenic
atoms, i.e. one has R = N D . This corresponds to the approximate propor-
tionality between cr and N o observed in Figures 1.16 and 1.17. If the carrier
concentration has Ihe value NLI>which at sufficiently high temperatures T i s
actually the caw, then the conductivity becomes independent of T. The only
remaining source of a temperature dependence for u is the T-dependence of
the mobility p. It is this relatively weak temperature effect which shows up
44

Figure 1.23: Ordering of the allowed en-


ergy values of electrons in a silicon crys- t
In
tal doped with arsonic (schematically). a,
Each arsenic atom correspond# to 1 lo- P
W
cslized energy level. G Eg
a
W

2
-
I

0 L
X--c

in Figure 1.17 at high T .


With this, the experimental observations relating to carrier transport in
semiconductors are explained microscopically, at least in principle. We now
proceed to the Hall effect.

T-0 T.0

0 L
X- X-

Figure 1.24; Distribution of the electrons of a silicon crystal doped with arsenic
over the allowed energy values. 011the left for T 0 K , OIL the right for T > 0 K .
7
45

*)U1-Phosphorus
doped Si
u
Boron
doped Si

Current 4 Current- a)

Current - Current- bl

Phosphorus
doped Si

~ a i current
i H ~ Icurrent
I + Cl

HOII current + fie'd Hall current d)

Figurc 1.25: nlrlstration of the Hall effect in two silicon smrples, one doped with
phosphorus, and the other with boron.

1.4.5 The Hall effect and the existence of positively charged


freely mobile carriers

We consider two semiconductor samples of the extrinsic type hoth madf of


the Same material, however, with different types of doping. To be specific,
we assume two silicon samples, one doped with phosphorus as a group V
dement. and the other with bnroii as a group 111 element (see Figiiie 1.25).
Applying a voltage to both samples, as shown in Figure 1.25a and b. a
current I will Bow. For both samples it has the same direction, namely from
'+' to '-', Consider, ROW, the ~ I I I Ptwo sarriples in a niagnPtic Geld B ( ~ e e
Figure 1 . 2 5 ~and d). As is well-known, the Hall effect will be observed in
such circumstances, i.e. a current component I H perpendicular to both the
electric field E and the magneliu field B will arise. Gnder the wndilforts
46 Chapter 1. Characterization of semiconductors

of Figure 1.25 no current can flow in this direction, and therefore there will
be a voltage UH such that the current caused by it will just compensate
the Hall current. Experimentally one finds that the Hall voltage UH has
different polarities in the two samples. This means that the corresponding
Hall currents flow in opposite directions. If we assume that the magnetic
field is directed normal to the plane of the figure and that it points into this
plane, then the Hall current flows upwards in the boron-doped sample, and
downward in the phosphorus-doped one.
The Hall effect can also be measured in metals. In this case the direction
of the Hall current is the same as in the silicon sample doped with phospho-
How can the dlfemnt behauiior of thesiicon sample dupd with born
be understood?
The explanation is quite clear if one assumes that, in contrast to metals
and the phosphorus-doped silicon sample, the current in the boron-doped
sample is not carried by negative charge carriers but by positive ones. This
is illustrated in Figure 1.25d. A charge q , which moves with velocity v in a
magnetic field B,experiences the Lorentz force

F = 4- [ v x S ] . (1.26)
C

Since negative charges move to the left, and positive ones to the right, the
Lorentz force F, which depends directly on the sign of the charge, has, for
both charge signs the same direction, namely upwards. For the boron-doped
sample this means a Hall current directed upwards, but in the phosphorus-
doped sample the Hall current is downwards, since in this case the charge
carriers are negative. This is just the observed behavior, which means that
the assumption of positive freely mobile charge carriers is successful in ex-
plaining the unusual sign of the Hall current in boron-doped silicon. In this
way the experimental observations of the Hall effect reveal a remarkable gen-
eral property of semicondiictors: For doping with certain atoms, the current
is not carried by negative charges as one would expect considering the nega-
tive charge of electrons, but by positive ones. In other words, in addition to
the electrons as negative freely mobile charge carriers, one also has positive
ones in semiconductors under certain conditions.
This surprising observation may be understood as ~OUQWS.One may
demonstrate - as we will do later explicitly - that boron atoms whi& sub-
stitute silicon atoms in tf silicon crystal, give rise to energy levels in the
forbidden zone just above the valence band edge. This is illustrated in Fig-
ure 1.26. Electrons in such energy levels are localized at the sites of the
boron atoms. At very low temperatures these states are not occupied, and
the boron atonis are electrically neutral. At finite temperatures electrons
from the valence band are excited into the boron levels (Figure 1.27), leav-
ing behind unoccupied states, or occupation hole3 in the valence band. We
1.4. Macroscopic properties and their microscopic implications 47

Figure 1.26: Ordering of the allowed en-


ergy values of electrons in a silicon crys-
tal doped with boron (schematically).
Each boron atom corresponds to 1 local-
t
i d energy level.

immediately recognize that these occupation holes of the valence band be-
have jubt like freely mobile positive charge carriers in an applied electric
field. This is illustrated in Figure 1.28. For simplicity only one hole is as-
sumed to be in an otherwisr rompletely occupied valence band, placed at the
upper band edge. The population of the bend edge is shown in Figure 1.28
at different points of time. At the beginning, the hole is at the outermost
left position. The adjacent electrons experience a force which tends to move
them to tho left. Rut only the electron neighboring the hole on its right
hand side can follow this force, since all other electrons are blocked because
the states to their left are already occupied. The electron immediately on
the right hand side of the hole, moves into the hole and leaves behind an-
other hole to the right of the first one. The hole has thus moved one site
further to the right in this way. 111 the next time interval this process is
repeated, and the hole again moves one site further to the right etc. Thus,
in an electric field, the hole moves like a freely mobile electron, but in the
opposite dirwtion, as if carrying a positive charge. In summary, holes in the
valence band behave like freely mobile positive charge carriers. This quali-
tative introduction of the concept of holes will later be elaborated by more
quantitative considerations.
As we have seen above, the sign of the Hall voltage tells one whether
the free carriers of an extrinsic semiconductor are negatively or positively
charged, Le. whether they are electrons or holes. In the first case one speaks
of a n-typr semzcondurtov (the n stands for negative), in the second of a
p-type semarondector ( p for positive). The antranszc semiconductors intro-
48 Chapter 1. Characterization d semiconductors

L 0 1
X- X-

Figure 1.27: Distribution of elect.rons of a silicon crystal doped with boron over
allowed energy values.

Figure 1.28: Empty states (halefi) in Ihe valence


band behaye like freely mobile poriitive charge carri-
em in an electric field.

duced earlier haw neither electrons nor holes from impurity atoms. Their
mobile electrons arr generated by thermal excitation of bound electrons from
the valence band to the rondiiction band. Since each excited electron leaves
behind a hole in the valence bmd. one also has mobile holes in intrinsic semi-
conductors. Their number is equal to that of the mobile electrons. Holes
are also present in n-type semiconductors, but, in very small numbers com-
pared to elwtrons. Analogously, a few electrons occur in p t y p e materials
with mob1 of the carriers being holes. One calls the many mobile carriers
of extrinsic semiconductors m a j o n l y cariwrs and the few mobile carriers
mzrrortty earners.
The Hall effect can also be used for purposes other than the determi-
nation of whethrr an extrinsic semiconductor is n- or p-type. The absolute
valup of the Hall voltage determines the concentration of majority carriers.
We will veTify this assertion for an n - t y p ~semicondiictor. In this case the
rurrcnt density j can be expressed in the form j =- --env. which permits us
l o rewrite equation (1.26) as

1
F=-[j X B ] . (1.27)
RC

This is the same force caused by the electric field E ~ I= F/(-e). The
voltage corresponding t,o EH is by dehition the Hall voltage Ufi. With b
as the width of the sample normal to j and 3,the expression below for U x
follows:

(1.28)
where
1
Rff = - (1.29)
PrtC

is the so-called Hall constant. We can write an analogous expression for


holes, only the electron concentration has to be replaced by the hole con-
centration p . Thin;, hy measuring thc Hall vultagr, onr ran also det.ermine
the majority carrier concentration.

1.4.6 Semiconductors far from thermodynamic equilibrium

All properties considered hitherto were for semiconductors in a state of ther-


modynamic equilibrium or in close proximity. However. semiconductors may
easily be driven into states far from equilibrium. Here, 'far' means that char-
acteristic macroscopic propexties of the semiconductor deviate strongly from
those in equilibrium.
To be more specific about these qualitative statements, we examine the
example of photo-conduction. Consider a sample of an intrinsic semironduc-
tor, shielded against unwanted influence of light. If, in this 'dark state' the
conductivity is measured. one obtains a relatively small value. in accordance
with the relatively low carrier concentration of an intrinsic semiconductor in
thermodynamic equilibrium. However, if one irradiates the sample with light
which is absorbed by the semiconductor (see Figure 1.22). the conductivity
will rise more or less strongly, depending on the intensity of the light and
the magnitude of the absorption coefficient. This is shown schematically in
Figure 1.29. assuming realistic conditions. When the exposure of the sample
t o light ceases, its conductivity decreases to the original low dark value. Ev-
idently, electrons in the conduction band and holes in the valence band were
created by irradiation with light to such an extent that their equilibrium val-
ues were exceeded by orders of magnitude. A simple estimate for Figure 1.29
50 Chapter 1. Characterization of semiconductors

Figiire 1.29: Conductivity of an intrin-


sic semiconductor irradiated with light
as a function of intensity I of the ab- 1M - -~
sorbed radiation (schematically).

I I
1~-7 - I
10-4 10-3 lo-7 10-1 ma r]'
I h o t t om* --+

shows that t.he carrier concentration was increased from about 10'' cm4
(equilibrium value) up to ~ light of intensity 10 Wcm-'. The
c ~ n . -by
electron-hole pairs created by the radiation decay, however, after a short
t,ime. 'lhis follows from the fact that continuous irradiation leads to tt sta-
lionary conductivity value, and lienre a constant carricr concentration, and,
on the other hand, from the observat.ionthat the conducbjvity decays down
to the dark value after switching off the light source. The 1at.ter observa-
t.ion means that, thermodynamic equilibrium i s reestablished by so-called
wxnnbinaiion, p7nuesses.
Rwidr irradiation with light, extreme nun-equilibrium states in semicon-
ductors can also be created in other ways, for example, by putting an n-type
semiconductor in contact with a p t y p e semiconductor or with a metal, or
by applying voltage to a. semiconductor which previously had been isolated
by a thin insulating layer from one of the electrodes. The ability to c r e
at,? extreme non-equilibrium states in semiconductors i s extensively used in
electsonic devices. Almost all applicat.ionsof semiconductors in such devices
rest on this uniquc pQssibility-.
Non-equilibrium processes in semiconductors and the most important
scmiconductor devices based o n them, such as electric rect.%ers, bipolar and
unipolar transistors, tunnel diodes, photodetectmu: solar cells, as well as
luminescence and laser diodes, will be dealt with in the second part of this
book, i.e. in Chapters 5 , 6 and 7. In the first part of the book, the basic
concept.s, discussed above in a heuristic way, will be developed from first
principles. This applies to the stationary electron states of an ideal semi-
conductor (Chapt,er 2), their niodifications by impurity atoms and other
deviations from the ideal crystal, as well as by external fields (Chapter 3)
and t h e statistical distribution of charge carriers over aiergy levels in ther-
modynamic equilibrium (Chapter 4).
51

Chapter 2

Electronic structure of ideal


crystals

In sections 1.2 and 1.3 of the preceding chapter we discussed the spatial
order and chemical nature of the atoms in ideal semiconductor crystals. The
present chapter is focused on the quantum mechanical energy eigenvalues
and corresponding eigenstates of electrons in such crystals. The substance
of this subject is summed up under the designation electronic structure. We
will show how the simple model of electronic structure of a crystal, the
energy band model, which we heuristically introduced in section 1.4, can be
rigorously deduced from the Schrodinger equation. The periodic ordering
of the atoms of a crystal in space is crucial for this proof. It will also be
seen that the eigenstates of electrons in a crystal devolve upon the electron
states of the free unbound atoms. This is immediately understandable if
one imagines that the crystal is grown from the gas phase, i.e. by chemical
bonding of previously isolated individual atoms. In this process the electron
states of the atoms will change, of course, but the resulting states, i.e. the
electron states of the crystal will also depend on the initial electron states
of the isolated atoms prior to crystal formation.

2.1 Atomic cores and valence electrons


Qualitatively, the kind of changes contemplated may be characterized as
follows: There is no doubt that the electrons of the outer shells, i.e. the
valence electrons, will react most strongly in assembling the isolated atoms
of a crystal, for they are the primary agents which bind the atoms into the
crystal state (Table 2.1). Whether, and to what extent the electrons of the
inner shells also change their states, is harder to predict. One may suppose
that such inner shell changes will be comparatively slight, at least for those
inner shells which lie much lower energetically than the valence shells (Ta-
52 Chapter 2. Electronic structure of id& crystsls

Table 2.1: Characterization of atomic cores and valence electrons of main group
elements from which semiconducting mat,erialsmay be formed.

Atom Core Valenw electrons


Nucleus Core electrons

B 5+ 222p
C 6+ 18 2922p2
N 7f 2s22p3

A1 13+ 323p
St 14+ 132222p6 323p2
P 15- 3s23p3

Ga 3 1+ 424p
Ge 32- ls22s22p633s23p63d10 3s24p2
AS 3 3+ 4Ap3

ble 2.2). Because the electrons of these deep shells are localized so close to
tlirir rrspertive nuclei. they feel potential changes produced by siirrouud-
ing atoms as being almost uniform. Strictly speaking, this means that the
wavefunctions of these electrons are essentially unaltered. Their eneqy lev-
els shift, however, specifically by the change of the constant potential value
arross their localization region. The term d i d d a t P shtjts is iised for t h e s ~
shifts of the inner electron levels. By measuring these shifts m e can obtain
information about the chemical nature and geometric striicturc of the envi-
ronnient of an atom in B solid. In B way, the inner electrotla threby serve
as probes.

Here, we arc intermtd in the electronic strizrture of crystals, dnd in this


regard the pertinent feature is that the wavefunctions of the inner shcll plec-
trons 01 the atoms in the crystal undeigo only weak changes. This statement
is even better justifid for the wavefiinctions of the nucleons in the atomic nu-
clei, which remain practically unaffected. Furthermore, if the gown crystal
is expoard to certain extrrrial perturbations - heat. prrssure, elwtromagnetir
fields - the states of the electrons of the inner shells and those of IhenucIeoas
frequently do not change. For the e x h n a l perturbations which are of p r i m
interest in semiconductor physics, this is even generally true. Therefore, in
determining the elertronic stricture of sanironductor crystals and the in-
fluence of exterruel perturbations on them, the states of the inner electron
shells and those of thc nuclei, as a rule, can be assumcd to be the same as
those of the free atoms. This allows one to consider the atomic nuclei and
2.1. Atomic cores and valence eJectrons 53

Table 2.2: Energy levels -E, and -4 of the valence electrons and - E , of the shal-
louwt core electrons for some chemical elements which may occur in semiconducting
materials. All energies are given in el/. (Afcer Herm.an. and SkiIlman, 1963.)

Atom 3 C N 0 A1 Si P S Zn Ga

-tp 6.6 9.00 11.5 14.1 4.9 6.5 8.3 10.3 3.1 4.9

-t, 12.5 17.5 23.0 29.1 10.1 13.6 17.1 20.8 8.4 11.4

-ec 195 291 405 537 87.5 116 147.4 182 20.7 31.7

-cP 6.4 7.9 9.5 3.4 4.7 5.9 7.2 8.6 3.5 5.8

the inner shcll drrctruns jointly as subsystems of the crystal, whose intrrnal
structure i s of no further interest sirire it does nut change. The structure is,
so to speak, frozen. In this sense t.hc subsystcms composed of atomic iuiclei
and inner electrons are elementary building blocks of the crystal. One rekm
t,o them as atom.ic ~ 0 : ~ s .

Since the crystal aIso contains valence electrons as independent particles,


we arrivr at a yic:i.ure which is fundamental for the further analysis - the
picture of a cryst,al as system composed of at,omiccores and valence electrons.
Somctirnes one refers to this concept, as the frozen-core a p p ~ ~ ~ i i m In
ati~~~
Table 2.1 the division into core8 and valence electrons is indicated for some
elements from which semiconductor crystals arc made.

The frozen nature of the electron states of the cores of a crystal! and
their lack of rcsponse to external influences, generally prevails, but as always,
there arc exceptions. In heavy metals such like zinc, t.he inner &shells are,
energetically, rather close t,o the oubar valence shells (see Tables 2.1 and 2.2).
In this case the d-electrons significantly participate in chemical bonding and
can no longer be included in t.he core, which is iinchangeahle by definit,iou,
Moreowr, the inner shell eledrons of crystals can be excited by means of
electromagnetic radiation in thc far UV and X-ray region. This can also
occur by means of an electron beam In solid state nuclear reactions, ~ W R
the states of the nuclei of the crystal atoms change.
54 Chapter 2. Electronic structure of ideal crystals

2.2 The dynamical problem

2.2.1 Schrodinger equation for the interacting core and va-


lence electron system
In dealing with the atomic structure of crystals in Chapter 1, we found
that their atoms are not located at arbitrary positions but at well d e h d
locations, specifically at points which are consistent with the existence of
a lattice and a unit cell. In this, the atoms were assumed to be point-like
and the space between them was empty. The crystal itself was imagined to
extend to infinity. We now consider a more realistic model of a crystal. First,
we replace the point-like picture of atoms by introducing spatially extended
atomic cores and take their centers of gravity as the atom sites. Second, we
recognize that the centers of gravity can also move to other positions than
those prescribed for the ideal crystal. In this way, we also account for the
fact that the atomic cores in crystals can execute oscillations around their
equilibrium positions, and that only these equilibrium positions form an ideal
crystal. Thirdly, the space between the massive elements of the crystal, i.e.
the cores, is now no longer assumed to be empty, as was done before, but
we acknowledge that valence electrons are present there. The assumption
of infinite extension of the crystal which, of course, is also not exact, will
be addressed at a later stage. This assumption excludes effects due to the
existence of bounding surfaces. As far as the electrons are concerned, these
effects are treated in section 3.6. The atomic cores will be marked by an
integer subindex 3 , and the valence electrons by an integer subindex d. Both
should start with 1 and run upwards, reaching arbitrarily large values since
we are considering an infinite crystal.
By means of a conceptual device which we are about to introduce -
notwithstanding the infinite extent of the crystal - only finite sets of J cores
and N valence electrons need be considered. To understand this, we imag-
ine the infinite crystal to be divided into parallelepipeds of macroscopic size
in such a way, that their edges are parallel to the primitive lattice vec-
tors a,,az,W of the crystal. These edges are to be given by the vectors
Gal, Gaz, G a with G a large integer. Each of these parallelepipeds should
contain an equal number J of cores and N of electrons. One calls these
parallelepipeds periodicity regions. Of all possible motions of the particles of
the infinite crystal, we now select those particular ones for which the cores
and electrons in different periodicity regions have the same positions relative
to the origin of their own region, and also have the same speeds. In this way
the infinite crystal becomes a periodic continuation of one particular peri-
odicity region, and it suffices to describe the motion of the J cores and N
electrons of this particular region. If the periodicity regions are made s a -
2.2. The dynamical problem 55

/; +
0
\ 0

Figure 2.1: 13escription of the positiom of the atomic cores ( 0 ) and valence elm
trons {a) {left part) as well as the interactions between these particles (right part).

ciently large, they will encompass all types of motions of an infinite crystal
with desired accuracy. The concept of the periodicity region makes it possi-
ble to pass from the original infinite space problem of motion to a finite one
without thereby losing the translational symmetry of the infinite crystal.
We use Xj t o denote the center-of-mass coordinates of the j-th atomic
core, and xi for the position of the i-th electron, which is further assumed
to be point-like {see Figure 2.1)- The j-th core mass will be denoted by
Mj. Of course, there are only as many diferent values of M j as there are
chemically different types of atoms in the crystal, so most of the Mj-values
are identical. In the case of electrons we can omit the index i from their
masses since they have the common mass m. The momentum of the j-th
core is called P3,and that of the i-th electron pi, such that

We are interested in the motion of the interacting atomic cores and valence
electrons of the infinite crystal, which can only be adequately treated by
means of quantum mechanics. The state of the system is described by a
wavefunction @, which depends-on the coordinates xi of all electrons and
Xj of all atomic cores, as well as on the time t. Since we assume periodicity
of the motion with respect to a periodicity region, it suffices to consider @
as a function of the coordinates xi of the N-electrons and the coordinates
Xj of the J cores of only one periodicity region. The state of the particles
in the remaining periodicity regions can then be described by means of a
periodic continuation of this function, i.e. by means of the relation
56 Chapter 2. Electronic structure of ideal crystals

@(XI I Ga,, x2 1 ,%:C . . . , XN i G%, XI 1 C%, X2 t G&, . . . ,X J t G+, t )


with CY 1 , 2 , 3 .
~

We will now wt up the Hamiltonian 7-l of the system of the N-electrons


and J atomic cores of a periodicity region. We use Tc and reto denote the
kinetic energies of the atomic cores and of the electrons, respectively, and
we define V ( X ~xz,, . . . ,XN, X i , X2,. . . , XJ) to be the potential energy of
the system. The Harniltonian is the sum of the kinetic and potential energy
operators,

+ + V.
H =- Tc Te (2.3)
The kinetic energies Xc and T, ran be expressed in terms of the momenta
Pj and pa of the cores and electrons as fOllOW8:

(2.4)

The potential energy is due to three interactions (see Figure 2.1):


(1) lhe repulsive Coulomb interaction of the electrons with each other. The
corresponding potential encrgy is denoted by Vce. It depends only on the
coordinates of the electrons, as given by

(2) The interaction of the electrons with the atomic cores due to their mutu-
ally attractive Coulomb forces, and also due to (repulsive) forces of quantum
mechanical origin, which become effective if the valence electron wavefunc-
tions overlap the inner electron shells of the atomic cores. The electron-core
interaction potential energy V,, depends on the locations of both the elec-
trons and cores. With respect to the electrons, it is evidently additive, i.e.

vet = vec(X1, x2,. . . ,X N , x1,XZ,.


. . XJ)=
I vc (xz,. x1,x2,.. . ,XJ),
(2.6)
where Vc(xi, X I ,X2, . . . , XJ) is the potential energy of the i-th electron in
the field of all cores.
(3) The mutual interaction of cores, which at sufficiently large distances is
again of Coulomb type. If the distances become small, repulsive forces of
quantum mechanical origin also occur. The core-core interaction potential
energy will be denoted by V,. It depends only on the locations of the atomic
cores, i.e.
2.2. The dynamical problem 57

vcc VCr(X1, XL 9 XJ). (2.7)


Summing the three potential parts (2.5), (2.6) and (2.7), one gets the total
potential

v - v, + v,, + v,, (2.8)


which determines the dynamical problem of the crystal uniquely. This prob-
lem i s described by the time-dependent Schrodinger equation

i3
iFL-e = H6,
at
whose solution may be determined in terms of the eigeiivalue problem for
the Hamilt,onittn 7f, which is given by the time-independenl Schrodinger
equation

%IJ! = EQ. (2.10)


The normalization condition for the wavefunction with refereiicc to a pcri-
odicity region is

(@I*) d3X1. . . d 3 x J p q X 1 ,x2,. . . , XN,X I , xZ, . . . , xj,t ) l 2 = 1.


(2.11)
Attempts to solve this eigenvalue problem exactly are hopeless from the very
beginning, because it involves a macroscopic system, i.e. a system with about
10 electrons and a similar number of atomic cores, the motions of which
are mutually coupled in a rather complex way. One must therefore resort to
approximations. Such approximations must first provide the means to reduce
the gigantic number of electrons, and secondly, allow for a proper decoupling
of the electron and core motions. The second simplification is achieved by
the so-called adiabatic approximation, and the first by the one-particle
approximation. These two approximations will be elaborated below. We
begin with the adiabatic approximation, and in the course of the discussion
it will also become clear how the somewhat unexpected designation of the
latter arises.

2.2.2 Adiabatic approximat ion. Lattice dynamics


The adiabatic approximation (also known as the Born-Oppenheimer approx-
imation) is based on the fact that the mass of the atomic cores is many tens of
thousands of times larger than that of the electrons - in Si, e.g., 52 thousand
times, and in mercury 368 thousand times. In addition, it takes advantage of
the fact that in a crystal the kinetic energy of an atomic core is, on average,
58 Chapter 2. Electronic structure of ideal cry&&

smaller than that of a valence electron. 'l'his can be seen in the following
way: If the cores and valence electrons were fwe particles, i.e. if they did not
interact, then the average kinetic e n p r o of a core would be approximately
(3/2) h7'. That of a valencp dzctron would be about (3/6) Eli where E F
is defined as the Fwmi energy of an electron gas of the same density. The
daerence between the average kinetic energies of the two types of particles
arises from the fact that the electrons obey Fermi statistics, whereas atomic
cores obey Boltzmann statistics. For typical concentrations of valence elec-
trons in a crystal of about 10" C W L - ~ the Frrmi e n e r a EF i~of the order
of m~gnit~ude of e V , while kT reaches only about 0.1 eV below the melt
mg point of the crystal. The average kinetic energy of a core is therefore
generally smdlcr than that of an electron. This remains true when the in-
teractions between the electrons and coresi which were omitted above, are
taken into account. Writing (M,/2) < X: > and (m/2) < x: > as the
avrragP kinetic enmgies of a core and an electron in a crystal, we thus have

Mj 1 7 1 . '
-2< X j > < -<xq>.
2
(2.12)
and it, follows that

(2.13)

Corrmpondingly, one may say that, on statistical average, the cores move
much slower than the electrons. This observation plays an important role in
the following considerations.
To simplify the notation, we replace the N-component sequence o the
vectors (XI,x2,.. . , X N )by x, and the J-component sequence of vectors
(X\,Xz,. . .,XJ)by X, ie. we write

x = (XI,xz,. , , XN), x = (XI,xz,. I . , XJ). (2.14)


To take advantage of the slow motion of the cores we write the solution
q{x,X) of the Schrodinger equation (2.10) for the total crystal. in the form
of a product

q x ,X ) = $(x, X ) ' @(XI. (2.15)


The necessary normalization (2.11) of the total wavefunction *(x, X) with
respect to a periodicity region is assured if each oI the two factors of (2.15)
is normalized with respect to this region, i.e. if

(2.16)
2.2. The dynarnical problem 59

(2.17)

are assumed. An analogous statement holds for the periodicity condition


(2.2) of @(x,X). To assure overall periodicity with respect to a periodicity
region we assume it for +(x,X) and 4(X)separately. An Ansatz of the
form (2.15) is always possible since it does not assert separability of the
variables x and X,but merely splits off a factor +(X)from the wavefunction
@(x,X)which depends only on one variable, X, while retaining the full
dependenre on both Coordinates in thr second factor +(x,X). The Ansatz
(2.15) becomes non-trivial if we proceed as follows: Firstly, we assume that
$(xiX)is the solution of a Schrodinger equation for electrons,

1% + v,,,c] X),
$(x,X)= U(X)@(X, (2.18)
where, for brevity, we have set

and U(X)is in the nature of an electron energy eigenvaliie. Secondly, we de-


mand that the split-off factor, 4(X), of the total wavefunction (2.15) satisfy
a Schrodinger equation in which the coordinates of electrons do not appear.
It turns out that such an equation cannot be derived rigorously, but only in
a special approximation - the adiabatic approximation which was mentioned
above. Yet without any approximation we have

[Te+ Tc + K , , + VCCI +(x,X)+(X)= E $ ( x , X)4(X). (2.20)

The set {$4) of the eigenfunctions of the Schriidinger equation (2.20) forms
an orthonormalized basis set in the Hilbert space of the crystal. Therefore,
relation (2.20) is satisfied if it holds for the Fourier type coefficients relative
to all basis functions Qd of this set, i.e., if the identity

($4lTe + Tc + v&ec + Vccl$,O)= E&!+t++ (2.21)


is valid. The necessary simplification concerns the matrix element ($4(Tc
I$4) of the kinetic energy of cores in this equation. Using relation (2.4)
between Tc and the squares Pj of core momenta, and applying the product
rule for differentiation we get, first of all,
60 Chapbrr 2. Electronic strncture of ideal crystals

The first two terms on the right-hand side of this q u a t i o n turn out to be
small compared to thr kinetic cuergy term of elwtrons in equation c2.20).
One has the order of magnitude relations

(2.23)

(2.24)

Here .n/E is a typird value of the core masses M,. The two equations (2.23)
and (2.24) are of fundamental importance in crystal dynamics, because they
are ultimately responsible for the drroupling of rloctron dynamics from the
dynamics of the cores. Therefore we present the proof of thwc equations
in Appendix B. Here we proceed a n the assumption that these relations are
proven.
The terms of rplativeorders of magnitude (rn/M)1/2 2 lov2 and ( m / h f )w
l W 4 , will be nPglatrd henceforth. With this the operator T, for the ram-
bined kinetic energies of all cores satisfies the approximate relation

Tc[.l$(x,X)9(X)l E=r ${x, X j W G q . (2.25)


This means that T, effectively does not act on V(X, X).In view of this rela-
tion we reconsider the SrhriidingFr equation (2.20) for the crystal, replaring
the terms which still depend on x by means of the electron Schrodinger equa-
tion (2.18) in terms involving the electron energy eigenvalue U ( X ) . Finally,
forming the scalar product with VI, we obtain the relation

(Tc + VC,(X)+ U ( X ) \ d ( X )= Ern(X). (2.26)


This rcpresents the SchrGdinger equation for the atomic cores in which the
coordinates x of thc electrons do not appear. The state of the elcctron
system, however, enters this equation, namely via its energy U(X)which
plays the role of a potential (referred to as adiabatze potential).
In summary, we h a w reached the following description of the total crys-
tal, viewed as an interact,ing system of atomic cores and electrqns:
The subsystemof electrons is describcd by R separate wavefunction qi(x, X),
which obeys a Schrijdinger equahion in which the coordinates of the cores
enter only as parameters in the potential, but do not occur as differential
operators in t,hr kinetic energy. In this way, the motion 01 the electrons is
treat,& as if the cores were at. red. Core motion, which does, of course,
occur despite its neglect with respect to electron motion, is described by
the wavefunction &(XIand the Schrodinger equation (2.26). T h e potential
of this q u a t i o n contains, besides the core-core interaction energy, a second
contribution IJ(X). This originsles in the interaction of the electrons with
2.2. The dynamical problem 61

the cores, for without such an interaction the eigenvalue U in the electron
Schrodinger equation (2.18) would be a constant independent of X, which
could be omitted. The potential contribution U(X) caused by the electron-
core interaction does not depend, however, on the electron coordinates. It
is an average value over all their positions X. The weight with which the
various positions x enter this average over the probability I + ( x ,X)1 d 3 N ~
of finding the electron system in a volume element d 3 N ~at the position x,
since equation (2.18) implies that

u(X) = (+(X)ITe + vee,ec(X)l+(X)). (2.27)


One can alternatively express this as follows: The electrons move so fast
that they are no longer seen by the cores as point-like particles, but as
smeared out over all space. Equation (2.26) for 4(X) thus contains the same
assumption as equation (2.18) for + ( x ,X),namely that the cores move much
slower than the electrons. In so far as this feature is seen from the point
of view of the electrons, equation (2.18) follows, whereas from the point of
view of the cores one obtains equation (2.26). Since the relation between the
velocities of the cores and the electrons, according to (2.13), is determined
by the inverse ratio of their masses, it is clear why this ratio must be small
for the two Schrodinger equations (2.18) and (2.26) to hold jointly in an
approximate sense.
It remains yet to clarlfy what effects are neglected because of the above
approximation and why this approximation is called adiabatic. In quan-
tum mechanics, one understands adiabatic temporal changes of potentials in
the sense that the changes proceed so slowly that no quantum mechanical
transitions will occur between the discrete quantum states of the potential,
which themselves evolve slowly from the initial onset of time variation. The
state of the system thus conforms continuously to the evolving new potential
values as a function of time, without any transitions to other states. That
exactly this situation is described by equation (2.18) and (2.26), may be
seen from (2.24). If one considers the previously neglected term in (2.24) of
relative order of magnitude ( m / M ) l I 2 ,then the total Hamiltonian 7-t of the
crystal has non-vanishing off-diagonal elements

(2.28)

and quantum mechanical transitions between the different eigenstates $14


and $+ of the crystal are recognized to occur. These transitions are caused
by the kinetic energy of the cores exclusively. If terms of the order of mag-
nitude ( m / M ) I 1 2 are omitted, then the quantum transitions due to core
motion are also neglected. This is equivalent to the assumption that the
core motion be adiabatically slow, in the quantum mechanical understand-
62 Chapter 2. Electronic structure of i d 4 crystals

ing of this term. The term adiabatic thus refers to the essential character
of the approximation in neglecting (rn/M)/. This approximation is use-
ful, of course, only as long as transitions between different eigenstates $$
play no important dynamical role. This is actually the case in regard to
many crystal properties and phenomena. There are, however, also effects
for which this does not hold, notably electric current transport. The fact
that the electric conductivity of an absolutely pure crystal does not become
excessively large is due in large part to the scattering of carriers from the
oscillations of the atomic cores, i.e. to non-adiabatic quantum transitions
between different electron and core states. Also, in the recombination of
electron-hole pairs mentioned in Chapter 1, these transitions play a decisive
role, with the lattice of atomic cores absorbing the energy which is released
during recombination. Formally, one may understand non-adiabatic transi-
tions as the result of an interaction between the electrons and the motion of
the atomic cores. Since such core motion, as we will see below, represents a
superposition of lattice oscillations, also known as phonon excitations, this
interaction is called the electTon-phonon interactaon
We have yet to explore how the two Schrodinger equations (2.18) and
(2.26) for the electrons and cores can be actually solved. The problem is
that both equations, are, at the outset, not completely determined - the one
for the cores contains the adiabatic potential U(X),which can be known
only after the equation for the electrons has been solved; and the electron
equation can be fully d e h e d , however, only if the positions X of cores in the
potential Vee,..-,-(x, X) are known. The direct way to overcome this difficulty
would be the following: One assumes a particular spatial ordering Xof the
cores and uses it to determine for them the potential Ve,ec(x,X).The lat-
ter is then used to solve the electron Schrodinger equation (2.18) (we will
not discuss here how this is accomplished, as it will be the subject of the
next subsection, 2.2.3). From the solution of the Schrodinger equation (2.18)
one obtains the value of the adiabatic potential U at the position X of the
cores. The same procedure is then applied to all other possible positions X,
whereby the adiabatic potential U(X) and the Schrodinger equation (2.26)
for the cores are completely determined. This equation can then be used to
calculate the core wavefunction t#(X). It follows that the dynamical problem
for the crystal as a whole is solved, since one would know its eigenfunctions
@(x, X)= y(x, X)q(X). In reality, however, this procedure is unsuccessful.
One cannot solve the electron Schrodinger equation for all possible core po-
sitions. Therefore, a simplified procedure is necessary. It contains additional
approximations, but has the advantage of being feasible in practical terms.
In this approach, one ignores the motion of the atomic cores completely and
assumes that they are resting in certain equilibrium positions Xq. In re-
ality, they execute oscillations around these positions with amplitudes that
become smaller as the temperature of the crystal decreases. However, due to
2.2. The dynamical problem 63

xo - - X" I.ke,ec(X,Xn)

? v,v,(xy -0 ?

no

J
xn+1

I
Figure 2.2: Iterative calculation of the equilibrium positions of the atomic cores.

the quantum mechanical phenomenon of zero-point oscillations, such motion


remains finite even at absolute zero temperature. The equilibrium positions
Xq are unknown at the outset. One can determine them by demanding that
the total energy

+
Vo(X) = U(X) VC(X) (2.29)
of the crystal in equilibrium have a minimum at Xq. Equivalent to this
is the requirement that the forces -VxVo(X) on the cores, the so-called
Hellman-Feynman forces, vanish at the equilibrium positions:

--vxvo(x)~x,xeq
= 0. (2.30)
Bearing this in mind, we may employ the iteration process below for the
solution of the two coupled adiabatic equations (see Figure 2.2): In this pro-
64 Chapter 2. Electronic structure of ideal crystah

cess, one assumes ccrtain trial equilibrium positions Xo, enters them in the
electron Schrodinger equation, and determines the eigerivalue U(Xo). This
solution is then used to determine the potential Vo(Xo) and the liellmaii-
F e y m m forces. Thanks to the Fe'egrman theorem, taken jointly with appro-
priate analytical transformations, one can determine these forces without
numerically calculating Lhe potentid in the environment of X'. After the
first iteration cycle, t.he Hcllman-Feynman forces will, in g~neral.not yet
vanish. signifying that thc cmes arc still not at equilibrium posibians. By
nieans of the non-vanishing f o r c e one det.ermines new trial positions XI.
'The new positions are then substituted again onto the electron Schrudinger
equation (z.18)! to calculate a new eigenvalue U(X1),and the latter deter-
mines the corresponding Heban-Feynrnan forces. This procedure is to be
repeated until the forces become zero. The corresponding core positions are
then the equilibrium positions XeQ.In this way one reaches a very impor-
tant result, the determination of the atomic structure of the crystal. Such
structure calculations are successfully carried out currently fw many solid
state systems, including a series of semiconductor cryst.als. With regard to
semiconductors, it can he shown, for instance, that under normal conditions,
Si haas the diamond structure, and that its Lattice constant a iR 5.49 A.

];at tice oscillations

In so far a6 atomic structure is concernedd:ouly the equilibrium positions of


the atomic cores are of interest. These can be understood as average values
(Q I X 14) of the core positions X with respect to the core wavcfmction 4 ,
However, the wavehnction 9 itself contains considerably more information.
It determines the probability distribution for the positions of the atomic
cores. That the probabilities of the cores being removed from their equi-
librium positions are non-zero is tbe quantum mechanical indication of the
existence of lattice oacdllatioru. These oscillations may be describd easily
using the S&rodinger eqriatiou (2.29) for the cores. In this: it suffices to
expand the potential I/o in a Taylor series with respect to hhe displacements
X - Xq from the equilibrium positions Xq,neglecting ternis beyond the
square term. The linear term of this expansion vanishes since the potential
energy has a minimum at Xq. Thus one obtains

VO(X) = vo(xeQ)i-' ( X - X e ~ ) V ~ v X r ~ ' o ( X-~x)e(4x1. (2.31)


2

With this potential the Schriidinger equation (2.26) for @(XIreads


2.2. The dynamical problem 65

where. for simplicity, Vo(Xeq) = 0 has been assumed. Equation (2.32) de-
scribes a system of coupled harmonic oscillators. The restoring forces are
determined by the second derivatives of the potential Vo(X). Using the
eigenvectors and eigenvalues of the matrix of restoring forces (actually, of the
so-called dynamical matrix which also includes the kinetic energy term), one
can easily transform to a system of uncoupled harmonic oscillators. Their
motions are called normal mode osciliataom or lattace oscillations,and their
excitation quanta are phonom.
Phonons are a good example of the introduction of a concept which is
of fundamental importance for the dynamics of many body systems, includ-
ing the many-electron system of a crystal which will engage us in the next
subsection. The concept we have in mind here is identified by the terms
elementary excztatzons or quast-particles (both terms are commonly used).
This concept is based on the possibility of decomposing the motion of a
system of mutually interacting particles - in our case of the atomic cores
of a crystal - into non-interacting components of motion - phonons in ow
case. The phonons or, more generally, the elementary excitations are. so
to speak, the elements of n o t z o n of the system, while the atomic cores or,
more generally, the actual particles. form the ~ ~ T U C ~ U Telements
-Q~ of the sys-
tem. ,413elementary excitation involves coordinated motion of all structural
elements of the system. Conversely, the motion of an individual structural
element is a superposition of all elementary excitations - the motion of the
atomic cores, for example, is a superposition of all normal mode oscillations
or phonons.
Besides the oneelectron and one-hole excitations, the phonons are the
most important elementary excitations, or quasi-particles, of a crystal. In
this book we will deal mainly with electronzc elementary excitations, and
will include phonons only if it is otherwise impossible to properly describe
electron dynamics. Relevant phonon information will simply be cited with-
out detailed justification, since a thorough development of the theory of
phonons is beyond the scope of this book. O w choice of subject matter here
is conditioned by the fact that electrons and holes are much more important
for understanding the properties of semiconductors as they are used in elec-
tronic devices, than are phonons. Readers who are particularly interested in
phonons are referred to other books (see, e.g., Born and Huang, 1968; Bilz
and Kress, 1979; Bonch-Bruevich and Kalashnikov. 1982).
We return now to the Schrodinger equation (2.18) for electrons. In the
sense of Figure 2.2, we approximate the core positions X by their equilibrium
values Xeq. As far as the latter are concerned, we take the point of view
that they are known from experimental structure investigations, e.g., by
means of X-ray diffraction. For common semiconductor crystals this is in fact
true in all cases. Taking this approach, the potential b=,- in the electron
Schriidinger equation (2.18) is well-defined from the very beginning. To
66 Chapter 2. Electronic structure of ideal crystals

simplify the notation, we suppress the core coordinates X in the potential


= V, + V, henceforth, writing Vec(x,X ) = v,(x). Similarly, WP write
the electron wavefunction @(x,X)as $:,(x). Usingequations (2.19), r1.5) and
+
(2.6), the Hamiltonian H = Te V e , , of the N-electron system in equation
(2.18 j takes the explicit form

2.2.3 One-particle appraximation. One-particle Schrodinger


equation
With the Hamiltonian H of (2.331, the N-electron Schrodinger equation
(2.18) can be written as

HVj,(Xl,x2,.. ., X N ) = Ull(X1, x2,.. . , XN). (z -34)


The wavefunction ~+4(XI, xp, . . . ,XN) must be periodic and normalized with
respect to a periodicity region. The Schrodinger equation (2.34) is impossible
to solve directly since it describes an interacting system of electrons having
a tremendously large number of particles of order loz2. The goal of this
~

subsection is to provide an approximate description which allows one to


reduce the number of particles down to minimum number 1. This will be
done by developing a oneparticle Schrijdinger equation whose solutions are
rdated to those of the true many electron Schrodinger equation in a well-
defined and sdiciently simple way.

Hartree approximation

In keeping with the remarks above, we a s s m e the existence of an infinite


set of one-particle wavefunctions q ~ l~, 2 . .: .. 'pm, from which the stationary
stat- $(xl,x2). . . ,XN)of the N-electron system may be constructed. The
p I I ( x ) ,v = 1,2,. . . , m , are, firstly, taken to be periodic with respect to
a periodicity region, as the wavefunction @(XI, x2,.. . , xN) itself, i.e. they
satisfy the condition

+
pv(x)= pu(x Cajj j = 1,Z, 3. (2.35)
Secondly, they are assumed to form a complete orthonormal set of functions
in Hillert space, sylribolically

(CPv~lPv) = 6uhr. (2.36)


Employing pY(x)we form wavefunctions for the N-electron system in the
folIowing way. We Brst associate each of the N electrons with a particular
2.2. The dynamical problem 67

oneparticle state pv[x), i.e. particle 1 with state pull particle 2 with state
puL,etc., up to partick N which is associated with the state. , ,oy Alterna-
tively, WP may say that we occupy state pw with particle 1, state p , with
particle 2, etc. Due to the Pauli exclusion principle, each state can host only
1 particle, ignoring spin (which we do at this stage). Thus a given state
p1, p2,. . . ,pm may occur among the papdated ones lpy, 'py, . , ,pVNnot
more than once. Most of the states will not even orcur o n r ~ i.a, , not at all
(bear in mind that there is an i n h i t e number of them). These states remain
unoccupied. The set of quantum numbers, (q, vz, . .. , VN),termed configzs-
ratzon, definea the state of the N-electron system uniquely if we understand
that the h s t number in this set refers to the state of particle 1, the second
to the state of particle 2 etc.. Henceforth, we abbreviate the configuration
( V l , V a l . . . ,w )by (.I.
Thirdly. we assign to each configuration (IJ)of the N-electron system a
wavefunction $(y)(xl,x2,. XN) which is given by the bllowing product of
I I

oneparticle wavefunctions:

${y}(xl,x21.. .XN)= 'pvl(XL)Lp65(X2).. . IP,(XN) =~LpvJx3)- (2.37)


j

Disregarding the miitual interaction of the electrons for the moment, the
product (2.37) forms an eigenstate of the N-electron system if the one-
particle wavefunctions pV,(xJ)are energy eigenstates of the individual o n e
electron subsystem Hamiltonians. This suggests t h e question whether a
similar result might be possible for interacting electrons, i.e. whether it will
be possible to choose the py,(xj) in such a way that the product statc +ivi
obeys the Schriidinger equation

H${V)(Xl* x2,.' ' I X N ) = ~ { V } ~ { Y } ( X lX72 r . ' . 1 XN) (2.38)


for the fully interacting N-electron system - if not rigorously. then at least
in some reasonable approximation.
To address this question one may use the variational principle of quantum
mechanics. In this procedureI the oneparticle wavefunctions pV,(xj) are
determined such that the expectation value of the N-electron Hamiltonian
H becomes a minimum for N-eIecctran states of the product form (2.37).
Here we take a slightly different approach, and start from the Schrodinger
equation (2.38). This procedure has the advantage that one ran ai. once
determine whether there is a suitable approximation in which d ~ { may ~ }
be written in the product form (2.37), and also barn the nature of that
approximation.
Considering an M-electron system, we label a particular electron i, and
this index can take all values between 1 and N. The Hilbert space of the
68 Chapter 2. Electronic structure of ideal crysWs

system of t.he N - 1remaining elect,rons 1,2,. . . , i - 1,i + 1,... ,N is spanned


by the set, of product functions

with P I , . . . , pz-l, p o + l . . . . , pjv ranging over all possible values 1,2,,


. . , co
independently of each other. In this remaining Hilbert space we form the
Fourier-type coefficients of the Schrodinger equation (2.381, i.e., we multi-
ply t,his equation by the complex conjugated product function (2.39) and
integrate over all XI,x2... . XN with the exception of xz. In this way we
obtain

(2.40)

Due to the orthogonality of the qv. t h e right-hand side of this equation


differs from zero only if the pvalues coincide with the v-values, i.e.. if
= q ,.., pz-l = ~ ~ - pcl,+l
1 , = v,+l. .... p~ = v w holds. However, the
left-hand side of equation (2.40) differs from zero if p J f v3 for one or sev-
eral I # t . Thus equation (2.38) cannot hold rigorously, which means that
the eigenfunctions ${,,I of H cannot be written exactly as a product {2.37)
of oneparticle wavefunctions. This is only possible under the condition that
the non-diagonal elements of the Hamiltonian operator on the left-hand side
of (2.40) may be neglected. It is this approximation which makes possible
the reduction of the N-particle wavefunctions to products of oneparticle
wavefunctions. It is called a one-partzcle approozamatton. Strictly speak-
ing, it is the simplest variant of a oneparticle approximation, the so-called
Hartwe appronmatzoa. .4 more accurate oneparticle approximation, called
the Hartree-Fock apprommatzoa, will be discussed below.
Within the framework of the Hartree approximation the equation system
(2.40) involves only the diagonal terms with p 3 = vj for each 1 # 2 , and
correspondingly takes the form

The diagonal elements in this equation can easily be evaluated as


2.2. The dynamical problem 69

(2.42)

On the right-hand side of this equation only the first three terms depend on
the electron coordinates x1 while the last two are constants in this regard. If
one substitutes the expression (2.42) into equation (2.41),then the last two
terms can be grouped together with U{'(.}to h m the new eigenvalur

Using the abbreviation

(2.44)

(2.45)

we rewrit,e (2.41) as

The final relation (2.46) has the form of a Schrodinger equation for the 7-th
particle where V'{(")(xp)is the potential energy of this particle and EV,is
its energy eigenvalue. Beside the potential energy ITc(&) due t o the atomic
cores. V'Iv}(x,) also contains the contribution \#'}(x,). It is caused by the
mutual interaction of electrons, and is commonly called the Hartree p o t e n t d
In explicit form, ITH%{.I (x,)reads

(2.47)
70 Chapter 2. Electronic structure of ideal crystals

Here the integration runs over a periodicity region. The Hartree putential
V~(X~) describes the potential energy of the i-tb particle in the Coulomb
potential produced by the charge distribution --e Ck+% i(oV,(x)l2of the re-
4
maining particles. The factor of the electron-electron interaction potential
(2.5) does not occur in expressions (2.45) and (2.47) for the Hartree poten-
tial since each electron pair contribntes only once. Obviously. the Hartree
potential and the corresponding energy cigenvalues depend OB the configu-
ration { v } of the N-electron system, and also on the index i of the particle
which was removed.
The one-particle Schrdinger equation (2.46) derived above for the i-th
elpctron, holds for each other electron as well. only with a somewhat different
potential. This difference will now be removed, together with the dependence
of the H a r t r e potential on the configuration {v} of the N-electron system.
We argue as follows: If the number A; of electrons is macroscopically large,
as in the case of the electron system of a crystal, and if we consider only
oneparticle states which are spatially spread out more or less evenly aver
the entire crystal, there is no signifkant difference if we extend the sum over
k in equation (2.47) for the potential V;,[(x,)t o include k - i . Then the
2-dependence of the potential no longer exists. The emor thereby incurred is
of relative order of magnitude l/N. If one considers, on the othcr hand, only
states ( v } of the &--electron system which are similar to each other, one may
also neglect the {v)-dependence of the potential and replace V{}(x) by the
value for a representative configuration {vo). The question is, does such a
representative codguration exist in the case of a semiconductor, and if so.
what is it. The answer to the former question is under normal conditions,
yes. For a representative configuration in the abovementioned sense, we
have the state of the N-electron system with lowest total energy, the so-
called ground state. In this state all one-particle states p with energies
Ev below a special energy value (the Fermi energy) are occupied, and the
states with energies above are empty. Under normal conditions the states
of the N-electron system which occur in semiconductors, and also in other
solids, deviate very little from the ground state, Non-normal conditions
are associated with large deviations, e.g., such as semiconductors which are
displaced to a highly excited state by intense laser irradiation. Excluding
such extreme cases, the Hartree potential Vrj}(x) for the configuration {v)
is almost the same as that for the ground state configuration {v}, and
correspondingly we have

(2.48)

For brevity, we set


2.2. The dynamlcal problem 71

V(X,) = V&) t VH(Xt). (2.50)


The extent to which the approximation of a Configuration independent Hartree
potential is valid again depends on the kind of one-particle states involved.
For the extended, planewavelike oneparticle states of an ideal crystal this
approximation works better than for the localized oncpartick states of a
real semiconductor. In the latter case the configuration dependence of the
potential may become essential (see Chapter 3 for further discussion).
Using (2.50), the Schrodinger equation (2.46) becomes

(2.51)

The Hamiltonian of this equation is the same for all particles and no longer
depends on the configuration of the N-particle system. Equation (2.51) is
therefore the oneparticle Schrodinger equation par excellence, devoid of any
reference to a particular particle or configuration of the N-electron system.
We may therefore omit the index i in equation (2.51). TJsing the oneparticle
Hamiltonian

P' (2.52)
ff = 2, I v(x),
this equation becomes

H'Pdx) = Evcpv(x). (2.53)


The Hamiltonian H of equation (2.52) is Hermitian, and it is natural l o
assume that its eigenfunctions form a complete orthonormal set in Hilbert
space. 'lhis assumption has in fact been made at the outset, with respect
to the oneparticle states cpu(x) forming the product wavefunctionu of the
N-electron system.
In summary, the discussion above has shown the following: Within the
framework of the oneparticle approximation, i.e. neglecting non-diagonal
elements of thc Hamiltonian, the product wavefunctions $ J { ~ } ( X x2,.
~, . . ,XN)
are eigenstates of the N-electron system provided that the oneparticle wave-
functions of the product functions satisfy the oneparticle Schrodinger equa-
tion (2.53). Solving this equation and forming thc product wavefunction
(2.37), one gets approximate solutions of the N-electron Schrodinger equa-
tion. In this way we have reached the goal which was formulated at the
outset to replace the N-electron problem by a oneparticle problem whose
72 Chapter 2. Electronic structure of ideal crystals

solution has a well defined and sufficiently simple connection with the solu-
tion of the N-electron problem.
The idea that the (py(x)are energy eigenstates and the E , are energies
of single electrons, underlying the above consideration, needs to be made
more precise. Because of the electron-electron interaction, the motion of a
particular electron is always tied to the motion of all others, and the energy
of an electron is also, in part, energy of interaction. The latter statement
manifests itself clearly in relation (2.43) between the one-particle energies E,
and the total energy U{.} of the N-electron system, which we will explore in
more detail. First of all, it can be further simplified. Using the one-particle
Schrodinger equation, one can re-express the terms on the right-hand side
of equation (2.43) by one-particle eigenvalues, leading to

The energy of the N-particle system is therefore not just the sum of all one-
particle energies. It is necessary to subtract the Coulomb interaction energy
of the particles. Therein is reflected the fact that the E, contain a certain
portion of interaction energy with other electrons. This is doubly counted in
the sum xi E , of one-particle energies, once in summing over the particles
themselves, and once in summing over their interaction partners, which is
done in E, automatically. To correct this, one must subtract the Coulomb
interaction energy.
This shows that the (p,(x) may be interpreted as states of single elec-
trons only in a generalized sense. In reality the (py(x)describe stationary
states of the motion of the N-electron system in which all electrons are in-
volved. These states of motion are not mutually coupled, as in the case of
normal oscillations of a system of interacting atomic cores. Using the ter-
minology introduced in that context one may consider the states q , ( x ) as
states of quasi-particles or elementary excitations of the N-electron system.
The E, are the corresponding quasi-particle or excitation energies. There
is, however, a qualitative difference between these elementary excitations
of the electron system and the normal oscillations of a crystal. This may
be made clear as follows. If one adds to the N-electron system (which we
will assume to be in the ground state) one more electron, i.e. if one passes
+
over to a ( N 1)-electron system, then the one-particle Hamiltonian (2.52)
does not change within the framework of the approximations made above.
The oneparticle wavefunctions pV of the N-electron system therefore also
approximately describe the elementary excitations of the system of ( N 1) +
+
electrons. This means that an eigenstate of the ( N 1)-electron system
may be realized by keeping the previously available N electrons in their one-
2.2. The dynamical problem 73

+
particle states and adding the ( N 1)-th electron in one of the oneparticle
states pu* of the N-electron system which were previously not occupied.
Thus, by adding an electron, the energy of the system rises approximately
by Ey*.This means that the eigenvalue Ev* of the one-particle Ilamiltonian
may be understood as the energy of an electron added to the system. This
statement is called Koopman h e o r e m . From it, one can learn more about the
kind of elementary excitations of the N-electron system that are described
by the p,. These are states in which, as always, all electrons of the system
are involved, but not all in the same way. Only one of the electrons is mov-
ing in such states, while the others play a passive role; they determine the
potential in which this movement occurs. One therefore refers to these states
as one-particle excitations of the N-electron system, and to their energies as
one-particle excitations energies or, in short, one-particle energies.
In addition to the one-particle excitations considered above there may yet
be others. This can be confirmed by taking a (N -1)-electron system instead
+
of the ( N 1)-electron system. The missing clectron corresponds to a hole
in a previously occupied oneparticle state ( p y ~ . The excitation energy of the
hole is -EvO, which did not occur among the oneparticle excitations consid-
ered above. It therefore represents an additional one-particle excitation. If
an electron is removed from state vy and simultaneously an electron is added
in statevT, then this corresponds to the excitation of the N-electron system
from state ($, Y:, . , . , v k ) into state (v;, v:, ..., I&). The energy difference
with respect to the ground state amounts to EV;- F 0 . It corresponds
5
to the excitation energy of a11 electron-hole-pair with the electron in state
pu; and the hole in state p 0 . If one excites a second electron from state
v1
'p o into state p,,;, the energy difference with respect to the ground state is
u!2
+
(Eq - Ey:) (By;- E e ) , etc. The excitation energies of the N-electron
system can thus be written as a linear superposition of oneparticle energics.
This is valid only within the one-particle approximation. In a strict seiisc
one also has many-particle excitations, which will be considered in more de-
tail below. As far as the one-particle excitations are concerned, there are no
others than the ones considered above, at least as long as one ignores spin
and the magnetic interaction between electrons.
It.is now appropriate to clarify how the oneparticle wavefunctions cpv(x)
111wbe calculated from the Schrodinger equation (2.53). The potential in
Uhis equation, more specifically the Hartree part VH(X), dcpends on the
wavcfunctions pV(x)which are involved in the construction of the ground
state of the N-electron system. One must know these functions in order to
write down the potential and thus define the oneparticle Schrodinger equa-
tion. On the other hand, these functions can only be obtained by solving this
equation. The situation is similar to that in the preceding section on the
coupled Schrodinger equations for the interacting system of electrons and
74 Chapter 2. Electronic structure of ideal crystals

d? - PU
n

+
V"(X) = t'c(x) VG(X)

Figure 2.3: Self-consistent solution of the oneparticle Schrijdinger equation.

atomic cores. As was done there. we may also solve the present problem it-
eratively (see Figure 2.3). We employ one-particle wavefunctions pf(x) close
to the true stationary oneelectron states. r s i n g cp:(x) we determine a PO-
tential t$(x) according to equation (2.491, form the total potential Vo(x) by
means of (2.50), and use this to solve the one-particle Schrodinger equation
(2.53). The solutions vt(x)are then substituted into formula (2.49), thereby
determining new potentials TG(x) and V l ( x ) . With the latter one recalcu-
lates the eigenfunctions p:(x) etc. One continues this iterative procedure
until the eigenfunctions, and with them also the potential in the follow-
ing iteration step. no longer change within a specified limit of accuracy. The
eigenfunctions and potential are then said to be determined self-consistently.

Spin a n d spin-orbit interaction

At this point in our treatment of the oneelectron approximation, it is a p


propriate to recognize that electrons h w e a spin, i.e. an internal angular
momentum with the two possible values rC/2 and --7i/2 in a given direction.
This is to say that electrons are capable of a motion in spin space, beside
2.2. The dynamical problem 75

their motion in coordinate space, which in this context is called orbatal m o -


tion. As orbital motion involves dependence of the wavefunction p ( ~on )
the space coordinate x,spin motion involves a dependence of p(x, s) on the
spin variable s which may take the two possible values s = and s = - z1
-
4
(below the latter value will be written as =_ -;). 'I'hus, in consideration of
spin, the wavefunction of an elcrtron changes from - an ordinary vector p(x)
i),
in Hilbert space to an element {~(x, p(x, ;)} in the product space of
the ordinary Hilbert spare and tht. two-component spin space, a so-called
two-component spinor. fiLnrtion, To determine the spinor state of an electron
uniquely, the quantum number X which defines the statr must also specify
the spin state. If the latter i s independent of the state in coordinate space,
this may be done by specifying another quantum number u for the spin mo-
tion along with the quantum number v of the orbital motion, i.e. by setting
- v , o where u may take-the two values T (spin up) and 1 (spin down).
The spinor {px(x, i), px(x, i)} can then be represented as a product of only
-
one spatially varying function py(x)and a spinor (~~(4)~
,yo($)} which does
not change in coordinate space. The two spirior components cpx(x, s) can
then be written as

(2.55)
In general, however, the orbital and spin motions are coupled. This is mainly
due to the fact that, on the one hand, the spin motion is accompanied by a
magnetic moment of the electron, and that, on the other hand, the orbital
motion gives rise to a magnetic field which couples that magnetic moment.
In quantum theory it i s shown that this interaction, which is called spin-ovhit
intemction, can be represented by the following additional term H,, in the
oneelectron Hamiltouian:
Tz
H,, - -[VV(X) x p] . (7. (2.56)
4m c
Here V(x) denotes, as before, the periodic crystal potential of equation
(2.50)) and 3 is the vector whose three components are Pauli's spin matrices.
In spin space one usually refers all quantities to the basis X I = (1,0),X I =
( 0 , l ) . Then the components of are

u.=(; i), u y - ( i0 - 0') 3


1 0
(2.57)

Taking account of spin and the spin-orbit interaction, the one-particle Schro-
dingw equation (2.55) in Hartree approximation t,akes the form
76 Chapter 2. Electronic structure of ideal crystals

(2.58)

Spin-orbit interaction i s in fact an important consideration in determination


of t,he energy specha of many serniconduct,ors.

Hartree-Fock approximation

An obvious drawback of the Hartree approximation is that the wavefunction


of the IV-particle system is not antisymmetric with respect to the exchange
of two particles, a requirements of the Pauli exclusion principle. 'l'his de-
fect can be easily remedied by replacing h e product wavefiinction (2.37)
of the Ilartree state by a linear combination of product wavefunctions with
exchaiigd partirle indices and altered signs. In conjunction with this, the
spin of the electrons has to be considered, such that the wavefunction of
, The antisymmetric linear
the t-th particle is given by the spinor p ~ , ( x ,si).
cornhatiom of the product waverunctions may he arit,tcn in t h e form of a
so-called Slater determinant

(2.59)

In this determinant, an exchange of t.he variables of two electrons leads to


the exchange of the two corresponding rows. The sign of the determinant
thereby changes, SO that the Slater determinant actually has the requisite
aiitisymmnctry propcrty. If t.wo of the quantum numbers XI, X2,. . . , AN are
q u a l , then two colunms 01bhP determinant are identical and vanishes.
'l'his means that no states of the N-electron system are allowed with two
electrons in the same oneparticle state. "his is just the Pauli principle,
automatically enforcd by the use of the det,erminant,alform of the N-part,icle
wawfunr tion.
Employing such SlaPer determinants as N-particle wavefunctions, as op-
p o s d to simple producls, a one-particle Schriidinger equation far pv(x) may
be derivd in the same way as before, but the potential in this equation ia
2.2. The dyynmiicd problem 77

somewhat different than that in the Hartree equation. It contains an ad-


ditional contribution, the so-called exchange potential Vay(x),and the total
potential reads

V(X) = K(x) + VH(4 + W X ) . (2.60)


In the case of negligibly small spin-orbit interaction, the orbital state may
be characterized by a separate quantum number v t , and the spin state by a
separate cpiantum number m i . T h c spinor components cp~,(x,, 8%) are of thr

+
form (2.55) in such circumstances. The sum VH(X) Vx(x) of the Hartree
and cxchange potentials can then be written in a relatively simple form. It
can be shown thilt theii action on the coordinate dependent factor of the
oncparticle wavefunction cpl,(x)takes the form

(2.61)

The Erst term on the right-hand side of this cquation is t,heHartree potential.
The factor of 2 results from summing over the t,wo spin states associated with
each wavefunction cpvk(x).The second term corrcsponds to the exchange po-
tential. Formally, it differs from the first term by exchanging the states at
the two positions x and x. The factor of $ reflects the fact that, firstly,
the exchange potential acts only between electrons of the same spin, and,
secondly, that for the ground state with total spin 0, half of the electrons are
in spin-down states, and half are in spin-up states. In this way the magni-
tude of the exchange potential is influenced by the existence of electron spin,
although its value is the same for spin-up and spin-down states. Equation
(2.61) also shows that, unlike the Hartree potential, the exchange potential
is non-local. The effect of the exchange potential on the wavefunction p,(x)
is represented by an integral operator. In actual calculations one often uses
a local approximation for Vx(x). The exchange potential proofs to be at-
tractive, which is to be expected: the anti-symmetric form (2.59) of the total
wavefunction ${A} means that the probability of finding two electrons with
the same spin at the same position is zero so that one has an exchange hole
around each electron. This lowering of electron density in the vicinity of an
electron results in an attractive potential in addition to the repulsive Hartree
potential since the total Hartree wavefunction (2.37) does not account for
the exchange hole.
The improved oneparticle Schrodinger equation with the potential of
(2.60) and (2.61) is called the Hartree-Fock equation, and the oneparticle
approximation, which underlies it, is called Hartvee-Fock approximation
Thereby, the Hartree and exchange potentials are understood as those for the
78 Chapter 2. Electronic structrrre of idea? crystah

ground state configuration v i of the N-electron system. The effects of the


electron-electron interaction, which are still neglected within the Hartree-
Fock approximation with configuration independent Hartree and exchange
potentials. are called correlation effects.

Correlation effects

Correlation effects are, first of all, manifested in the fact that the true one-
particle excitation energies of an N-electron system differ from those in the
Hartree-Fock approximation. In particular, these excitation energies depend
on the configuration of the system, one has tt, configuration dependence. Sec-
ondly, Slater determinants which in Hartree-Fock approximation are con-
sidered to be eigenstates of the total Hamiltonian, in fact do not diagonal-
ize tbis Harniltouian ezactlv; there are non-vanishing offdiagonal elements,
an effect which is termed configuration interaction. The exact eigenstates
of the total Hamiltonian are linear combinations of diflerent Slater deter-
minants, and the corresponding energy eigenvaluea are no longer s u m of
oneparticle excitation energies, as had been the case for individual Slater
determinants. In other terms, the exart,eigenalates of the N-elwtrronsystem
are not oneparticle excitations! but many-particle excitetians. Examples of
many-particle excitations include two-particle ezcitation.5 of an electron and
a hole which are bound together by their Coulomb Interaction. The exci-
tation energy of such a hound electron-hole pair, the so-called ezcitan, is
smaller than that of the excitation energy of a free electron and hole pair,
differing by the binding energy of the pair. The reason for the designnation
correlation effect for this phenomenon is obvious: binding may be under-
stood as a correlation between the positions of the electron and the hole,
since their separation by a distance of about a Bohr radius is more probable
than all others. This interpretation presents the correct concept of correla-
tion in other cases also the states of the electrons are no longer independent
~

of each other, but, are correlated contrary to the assumptions implicit in the
oneparticle approximation. Collective many-particle excitations are excita-
tions of states in which all electrons of the system participate in comparable
measure. Examples include the plasma o s c i l - t i a m of an electron system.
They form a direct electronic analogy to the lattice vibratious of the atomic
cores of a crystal. Their excitation quanta are called plasmons.
The consideration of correlation effects stands along the most difficult
problems of solid state thmry which, even today, is not completely solved.
A comprehensive analysis of this problem i s far beyond the scope of the
present book. Readers who are particularly interested in correlation effects
will find discussions in a number of textbooks (see, e . g , hbrikosov, Gorkw,
md Dzyaloshinski, 1963; Fetter and Walecka, 1971; Ziman, 1974; Callaway,
1976; Madelung, 1978; Harrison, 1981). Below we summarize some results
2.2. The dynamical problem 79

which will be needed in Chapter 3. In doing so, we will concentrate on one-


particle excitations, i.e. individual electrons and holes moving in the force
field of all other electrons as well as in the force field of the atomic cores.

Correlation effects on one-particle excitations. Density functional theory.

Correlation effects on one-particle excitations may be treated by means of


the Green's function theory of many particle-systems. The poles of the o n e
particle Green's function in the complex energy plane represent oneparticle
excitation energies (more strictly speaking, the real parts of these poles are
the energy levels, and the imaginary parts are the lifetime broadening en-
ergies of the one-particle excitations). The one-particle Green's function is
governed by the Dyson equation, which contains correlation effects through
the so-called mass or self-energy operator.
Another method which works well for oneparticle states involved in the
ground state of the many-particle system is known as density functional
theory. This method relies on a theorem, the Hohenberg-Kohn theorem,
which ensures that the ground state energy Eo of an interacting electron
system in an external potential Vc(x) is a functional E o [ n ( x ) ]of the total
electron density n(x) of the ground state alone. This implies, first of all, that
the total energy E o [ n ( x ) ]depends on the oneparticle wavefunctions only
through the ground state density n,(x)and, moreover, that the density enters
at every point x, through an integral over X . The oneparticle wavefunctions
determine the ground state density by means of the equation

(2.62)

where cpvi(x) denote the one-particle states which, in the ground state of
the N-electron system, are populated by electrons z = 1,2, ..., N. According
to the variational principle of quantum mechanics, the wavefunctions cpv,(x)
adjust so that, while keeping their norms (cpudlqv,) constant, the total en-
vrgy Eo[n(x)]is minimized. This requires the vanishing of the variational
-xi
derivative of the functional E o [ n ( z ) ] E,(cp,Ip,) with respect to p:*(x),
where the factors E , are variational parameters, therefore

(2.63)

In this functional derivative the value of cp:t(x) at a certain point x is taken


as an independent variable, with respect to which the common derivative is
taken.
The total energy functional Eo[n(x)]of the ground statr may be decom-
posed into several energy contributions, namely, the kinetic energy E ~ & ( x ) ] ,
the external potential energy E c [ n ( x ) ] the
, Hartree energy E w [ n ( x ) ]and the
80 Chapter 2. Electronic structure of ideal crystals

exchange and correlation energies which are usually summed in the exchange-
correlation energy E x c [ n ( x ) ] .Thus

+
Eo[n(x)]= Ekin[n(x)l Ec[n(x)l &dn(x)l + + Exc[n(x)l. (2.64)

The functionals E c [ n ( x ) ]and E ~ [ n ( x are


) ] easily obtained as

(2.65)

E ~ [ n ( x=
)]
2
//
R R
d3x'd3x
n(x') . n ( x )
Ix' - XI '
(2.66)

The functional E x c [ n ( x ) ]is less obvious. It is usually taken in a local ap-


proximation called the local density approximation (LDA). The LDA starts
with the homogeneous electron system without any external potential. In
this case the density n ( x ) is a constant n in space, and E x c ( n ) reduces to an
ordinary function of n. Dividing E x c ( n ) by the total number nR of electrons
yields the exchange-correlation energy E X C ( ~of) the free electron gas, per
electron. The total exchangecorrelation energy Exc(n) of a weakly inho-
mogeneous electron gas of density n ( x ) should then be given approximately
by the expression

(2.67)

Finally, the kinetic energy fuIictiona1 E:k.tn[n(x)]has to be specified. By


definition, we have

(2.68)

Although this expression does not look like a functional of n ( x ) il is in-


deed possible l o transform it into such a form because all other terms in the
total energy functional Eo[n(x)]of (2.64) are functionals of n ( x ) , and the
Hohenberg-Kohn theorem enforces this for E & t ( x ) ] . For the waluation of
the variational equation (2.63) we do not, however, need the explicit func-
tional form of EkznIn(x)l;expression (2.68) suffices. Its variational derivative
with respect to p;,(x) iu given by

(2.69)

For the rcmaining functional derivatives, it follows that

(2.70)
2.2. The dynamicd problem 81

(2.71)

(2.72)

where
(2.73)

denotes the ~xcliangPcnrrelationpotential. The latter can be determined


if the exchange-correlatiou energy E x c ( n )of the homogeneous electron gas
is known as fniictio~iof n. This dependence can hr obtained by calcrilat-
ing E>yc(n)numerically for tfifkrent values of n and then lilting the data
to appropriate explicit functions. TJsing this procedure. various Pxchang+
correlation potentials have been proposed, for exaniylr

Lkc(x)- - (:) 1/3


e2n1j3[x)[I + 0.7734 z In
with z - rS/2l, where rS = and CI.Bthe Bohr radius (Hedin,
Lundqvist, 1971). Suhstitiiting i n h q u a t i o n (2.63) t h e frlnctionrtl drriva-
tives obtained above, one arrives at

(2.75)

with
v(x) = VC(X) I Vrr(X) + i<Tc:(x) (2.76)
as an effective one-elcrtron potential. Th? electron ind~xa has heen omitted
here bwauw thP equation is the bame for all electrons.
Relation (2.75), with the potential V(x) given by (2.76), is known as
Kohn-Sham equatzon. ,4s rompnred t o the oneeleclron potential V(X) of
the Hartrw ur Hartret-Fock equations, that of the Kohn-Sham equation
additionally accounts or correlation effects. The physical significance of
tlir solutions of the Kohn-Sham equation is, however, less direct than that
of the solutions of thP llartree or Hartrw-Fock equations. Generally, the
eigcnvalucs of the Kohn-Sham equation cannot be understood in the sense
of oneparticle excitation energies of the N-electron system. as it is possible
for the eigenvalues of the IIartree or HartrePFock quations according to
Koapman's theoiern. A misinterpretation of this kind may lead to large
errors. This applies, in particular, to electron-hole excitation energies in
senlimnductor crystals, defining the energy gap. The resulting erroneous
gaps are about 50% smaller than the experimental values. However, the
82 Chapter 2. Electronic structure of ideal crystals

eigenvalues and eigenfunctions of the Kohn-Sham equation can be properly


used to calculate the total energy of the ground state of the N-electron
system by means of the total energy functional (2.64). The dependence of
the total energy on external parameters as, for example, on the positions of
atomic cores, also can be obtained in this way. Minimizing the total energy
with respect to the core positions yields the atomic structure of the crystal.
The oneparticle excitation energies may be obtained as the differences of the
total energies of the ( N t 1)-or ( N - 1)-electron systems, on the one hand,
+
and the N-electron system, on the other hand, where the ( N 1)-electron
system applies for electlan excitations and the ( N - 1)-electron system for
hole excitations.

2.3 General properties of stationary one-electron


states in a crystal
The oneelectron Schrodinger equation (2.53) is specified by the form of the
potential V(x) which is different for crystals of different chemical composi-
tion and atomic strurlure. However, there are certain general properties of
V(x) which do no1 depend on the particular material nature of the crystal.
As we already know, a crystal remains invariant under a transformation by
an element of i t s space group. Rotations, reflections and rotation-reflections
of the point group of directions transform a given crystal direction into a
physically equivalent one. The symmetry of the crystal is transferred di-
rectly to the potential V(x). Consequently, the Schrsdinger equation (2.53)
is endowed with corresponding symmetry properties. We shall first describe
these and then explore their implications for the stationary oneelectron
states cpu(x). In doing so, we initially neglect electron spin.

2.3.1 Symmetry properties of the one-electron Schrodinger


equation
For simplicity, we restrict our considerations here to crystals their space
groups are symmorphic. These are groups which contain solely translations,
rotations, reflections as well as rotation-reflections, while screw rotations
and glide-reflections are excluded. The general case of non-symmorphic
space groups, and with it the particularly iniyortant case of the diamond
structure, i s treated in Appendix A. It turns out that the results derived
for symmorphic space groups arc also valid, with minor modifications, for
non-symmorphic ones.
We use the symbol t R to denote the lattice translation operator which
causes a translation of all points x through a lattice vector R,

tRx =x + R. (2.77)
2.3. General properties o f stationary oneelectron states in B crystal 83

The set of all lattice translation operators forms the translation group, which
we already encountered in Chapter 1. Since the lattice translation operators
are symmetry dements of the crystal, the translation gioup is B subgroup of
the space group which, by definition: contains all symmetry elements of the
crystal. As in the case of translations, we also assign operators to rotations,
reflections and rotation-reflmtions of the point group of directions. These
operators also art on the spatial vectors x, and W P denote them by the
symbol a and call them point symmetry operataorw. For the symrnorphic
space groups considered here, each element y may be thought of as a product
0 . t or r a translcttion t~ and a point syrnrnetrg operation a. Since an
~ t ~ r of
arbitrary position vector x may be represented in a basis spanned by the
three primitive lattice vectors al.a 2 and a3, it suffices to s p e d y the &ect
of a on these. We define (see Appendix A}

(2.78)

with ov as real coefficimts. By means of this relation each opeiator a is


uniquely associakd with a corresponding matrix aV. If the thee primitive
lattice vectors a, are orthogonal and of the same length, the uaJform orthog-
onal niatriw$, i.e. their inverse matrices are the same as their transposed
ones. However, this holds only for the primitive cubic lattice, and is not true
fur all other 13 Binvais btticrs. Thus, the utJ arc not iu general orthogoual
matrices. The effect of a on an arbitrary position vector x may easily be
d e t e r m i n d by ~ n e a n sof i t s decomposition

x -czja3 (2.79)
t

with respect to the three primitive lattice vectors aj. The xj are the com-
ponents of x with respect to these vectors. Applying a ,

(2.80)
j

we may rewrite this relation in the form

(2.81)

from which it follows that the transformation a , which was ariginaIly defined
as a transformation of the basis vectors ai with fixed coordinates zz,may also
be understood as a counter-tra~sformation of thc coordinates with fixed basis
vectors. Indeecl, lh coordinate counter-transformation takes place with the
84 Chapter 2. Electronic structure of ideal crystals

transpose matrix a j i of aij, in contradistinction to the transformation (2.78)


of the basis vectors. This is t o say that

(2.82)

If the crystal remains invarianl under the transformations t R and a , then


this mist also hold for the potential V(x) with which the tramformed crystal
acts on an electron at position x. To express this fact formally, we &fine
the operation of t~ and n on an arbitrary position dependent scab1 crystal
function S(x) as follows:

tf$(x) = S ( t R ' X ) , (2.83)

a S ( x )= S ( a - l x ) . (2.84)
It is striking that the operators t~ and a act on S(x) in such a way that
x ifireplaced by tR1x or m - l x . but not by k ~ orx c k x . as one might hme
expectd. The rhosen deftnition strms from the rrcognition that the trans-
formed property is that of the transfarmrcl crystal at the original position
x. This is the Sam?, howewr. as the proprrty o f t h e original crystal at
the inverse transformd position. Formally the definitions (2.83) and (2.84)
giiarantee the correct multiplication order of two no11 commuting operators
2 . Applying these
under the fiinction symbol, because ( q c q ) - ' - rtl-1 a -1
definitions to the potential V ( x ) and simultaneously requiring crystal sym-
metry, w r have

tnv(x) = V(t,lX) - v(x), (2.85)


UV(X)= V ( d X ) = v(xj. (2.86)
Further application of these opwatm relations to a wavefunction p(x) one
obtains

and since ~ ( xmay


) be chosen arbitrarily, the relations

(2.89)
(2.W)
follow, with IA, B ] = A B - B A as abbreviation for the commutator of two
operators A and 3. One says that t~ and 1y commute with the potential
V ( x ) . Such cornmutivity also holds for the uthcr contribution of the Hamil-
tonian, the kinetic enerw operator ?' = p2/2m. For t~ this follows directly
2.3. General properties of s i a i h n a r y oneelectron s&atwin a crystal 85

from the relations p = - i h V , and V, = V x + ~while


, for a it follows from
the fact that p2 is the square of the length of the momentum vector, as
well as from the observation that the length of a vector is not changed by a
translation, rotation or reflection. Thus.

2-1 = 0.
[tR,TI = [a, (2.91)
Taken together with ('2.89) and (2.90), these relations yield

[ t R ,HI = 0, (2.92)

H ] = 0.
]u, (2.93)
Since the elements g of a symmorphic space group may be written as prod-
ucts of t~ and C Y , one also has

The latter relation expresses in h a 1 form the implication of crystal symmetry


for the Schrodinger equation. We will use it extensively in the analysis of
symmetry properties of stationary oneelectron st,ates to follow.

2.3.2 Bloch theorem


Let p~ be an cigenfunction of H having eigenvalue E . Then the stationary
Schrodinger equation (2.53)holds, whrrc the state index v has been replaced
by E .
HVEW = EPE(4. (2.95)
In addition, one has the periodicity condition with respect to a periodicity
region
VE(X -k G a j ) = ( P E ( X ) 1 J - 2, 3, (2.96)
and t h e normalization condition

Through the periodicity condition (2.961, the symmetry group, which at


the outset includes an infinite number of lattice translations, is reduccd to
the finite subgraup containing only those translations t R which do not fall
outside the periodicity region. The commutivity of t~ and a with H has the
consequence that, along with cp~(x), also t ~ p ~ ( xw,uE(x)
), and g p ~ ( x are
)
eigenfunctions of H having ihe same cigenvalue E . Thus, one has
86 Chapter 2. Electronic structure of id& crystah

(2.100)
Consider now the eigcnfunction g i p p If the symmetry operation g ranges
over the whole space g o u p , the totality of vectors g p spans ~ a subspace of
the Hilbert space w l i o s ~dimension d i s in general larger than 1. This is to
say that the eignvalue E, for symmetry reasons, is d-fold degenerate. It
can be shown that, apart from spmial cases, d equals the number of dif-
ferent elements m of the point group, and then the d basis functions of the
subspace may be chosen in the form with Q ranging over the entire
point group. This result will not. howevpr, be usrd in what follows and the
basis funrtions will be written in the general form p ~ lp, ~ 2 ,. .. , IpE& They
span that subspace of the Hilbert space which contains the eigenfunctions
of the Hamiltonian with the eigenvalue E . According to i t s construction
this space is invariant under the operation of an arbitrary element g of
the space group. In terms of the concept of rmducible represenfatema of
groups (an introduction is provided in Appendix A), we may alternatively
express this observation by saying that the eigenfunctions of H for a partic-
ular eigenvalue E give rise to a &dimensional irreducible representation of
the space group. The same statement also follows for each subgroup of this
group, particularly for the subgroup of all translations. The space spanned
by r p ~ lP, E T , . . . , ( p ~ d :thus also provides B representation of the translation
group. However, this representation is, in general, no longer irreducible.
That means that the original basis p ~ pl ,~ 2 .,. . ,( P E may ~ be transloorrued
(d-1) x I
into a new basis (p~,(pk,'plh;,.,pE in such a way that the represen-
tation matrices of all translation operators t~ written in the new basis are
constructed from lower dimensional matIices, odered along the diagonal. It
is of special importance that the translation group is a group of Abehan type:
The resiilt of two translations t~~ and t~~ does not depend on the sequence
in which they are executed. Formally this is expressed by the equation
(2,101)
As is demonstrated in Appendix A, all irreducible representations of Abelian
groups are 1 dimensional. This means that the lower dimensional matri-
ces along the diagonals of the tH-representation matrices are 1-dimensional.
(d-1) x I
Each of the basis functions &, ,p;c . . . , ' p E therefore forms a repre-
srtntation space of the translation group by itself. Thus, su functionstRpE,
with t~ an arbitrary element of the translation group, are linearly dependent
on p~ and dl functions t ~ &are linearly dependent on & etc. 'I'his means
that thp t p p itself
~ may be written in the form
tRcPE(x) = c(R)rFE(x) (2.102)
where r(R) is a complex coefficient. Analogous equations hold for all o t h a
functions t ~ l p btR&.,
, , . .. This is equivalent to the statement that the cho-
2.3. General properties of stationary oneelectron states in a crystal 87

sen ( P E ,'pk,&, . . . are eigenfunctions of the translation operator. This result


forms the content of the
Bloch theorem:
The eigenfunctions p~ of the Hamiltonian H of a crystal can be chosen
such that they are simultaneously eigenfunctions of the lattice translation
operators of the crystal
The particular energy eigenfunctions whose existence is stated by this the-
orem are termed Bloch functions. The proof of the Bloch theorem sketched
above relies in an essential way on features of group representations. Al-
though it forms the most appropriate proof, one may, however, also proceed
without the tools of group theory, using a mathematical theorem which eo-
s u e s that two commuting Hermitian or unitary operators have a common
set of eigenfunctions.
The Bloch theorem is of such great importance that several remarks are
appropriate. In the first place, it has to be emphasized that this theorem is
an immediate consequence of the commutivity of translations. For the sym-
metry operations a of the point group, for example, which in general do not
commute, it does not hold, which means that in general the eigenfunctions
of H cannot be chosen to be simultaneous eigenfunctions of the operators u
of the point group. Secondly, the theorem does not say that every conceiv-
able eigenfunction of W is also necessarily an eigenfunction of t R . This holds
only for specially chosen c p ~ .The Bloch theorem insures that such a choice
is always possible. Thirdly, this theorem also does not imply that only one
eigenfunction exists for particular eigenvalues E and c(R) of, respectively, H
and t R . In reality there are alwtLys several. The pair of eigenvalues El c ( R )
is therefore not sufficient to uniquely characterize the eigenstates of H and
of the group of operators t R . Quantities which provide such unique charac-
terization have yet to be identified. It, turns out that there are three real
numbers k l , E z , k3 which allow one to distinguish the different irreducible
representations of the translation group of the crystal.
To understand this we must examine these representations in greater
detail. They are defined by the eigenvalue equation (2.102) of the translation
operators. T o start, we will show that the eigenvalues c k l k 2 k 3 ( R ) of this
equation may be written in the form

(2.103)

P k 1 k 2 k 3 ( R )= (-2r)(hri + ~ Y+ zh ' 3 ) . (2.104)


Here, the E l , kz, k3 are the above mentioned real numbers which determine
the representations uniquely. The factor L 2 ~is1 introduced to simplify ex-
pressions which later will arise. To prove equation (2.103) we first show
88 Chapter 2. Electronic structure of ideal crystals

(2.105)
(the subscript indices k l , k2, Ic3 on c(R) and P(R) will be suppressed tem-
porarily). The proof is based on the normalization of p~ according to equa-
tion (2.97), which leads to

I I
(tR(PE t R P E ) = ( P E (PE). (2.106)

This holds because the translation t~ of the integration variable through R


in ( t ~ pI t ~ p may
) be absorbed by a change of variables jointly with an
application of the periodicity of the wavefunction p(x) with respect to the
periodicity region. On the other hand, it follows from relation (2.102) that

I
(tRLPE tR(PE) -I c(R) l2 (PaE I P E ) . (2.107)
Considering (2.106) and (2.107) together, (2.103) is verified at once. Sec-
ondly, we show that

c(R1+ R2) = c(R1) . c(R2) (2.108)

must hold. To prove this, we use the following obvious relations:

Comparison of the last two relations immediately shows that (2.108) is also
true. Employing (2.103) for c(R), equation (2.108) now yields

. ~xP[WWI,
exp[iP(Rl+ Rdl = exp[iP(R~)I (2.112)
whence
P(Ri t R2) = P(R.1) t P(R2). (2.113)
This means that P(R) is a homogeneous linear function of the components
TI, 7-2, r3 of R. As such P ( R ) must have the form (2.104).
We now proceed to the eigenfunctions of the translation operator. Em-
ploying ( ~ ~ k(x) ~ to
k denote
~ k ~ the eigenfunction having eigenvalue Ck1kzk3 ( R )
of equation (2.103), we claim that pEklkzks(x) can be written in the form

(2.114)

where u ~ k ~ k ~ k denotes
~ ( x ) a lattice-periodic function, such that, for any
lattice vector R,
2.3. General properties of stationary oneelectron states in a crystal 89

UEklkzk3(X - R) = UEk1kzk3(X)' (2.115)


The factor l/& with 0 = G300is introduced in (2.114) in order that the
normalization integral of UEklkzk3 (x) with respect to a primitive unit cell be
1. The proof of (2.114) may be carried out by verifying that functions of the
form (2.114) obey equation (2.102). This results in

which we know to he true. The functions p ~ k ~ k ~ k of ~ ((2.114)


x ) are referred
to as Bloch functions and the factors u ~ k ~ k + ~ (as
x )Bloch factors. Recalling
that the p ~ k ~ k ~ are
k ~also
( ~ simultaneously
) eigenfunctions of the Hamilto-
nian, we may express the Bloch theorem in a somewhat more specific form
than we did above:
The ezgenfinctions of the one-electron Hamiltonaan of a crystal can be chosen
i n the f o r m (2.114).
The real numbers kl,k2, k3 characterizing the various Bloch functions may
he understood as components of a vector. However, as we will soon see, this
is not a vector in coordinate space, but one in a space which is reciprocal to
coordinate space.

2.3.3 Reciprocal vector space and the reciprocal lattice


The starting point for understanding the nature of the components k l , k2, k3
is their transformation law under point symmetry transformations a in co-
ordinate space. This will now be explored.

Transformation properties of kl,122, k3

At the outset, it is clear from equation (2.99) that both p ~ k ~ k ~ k and ~ ( x )


a q E k l k z k 3 (X) are degenerate eigenfunctions of the Hamiltonian H with the
same eigenvalue E . The question arises whether a p , g k l k Z k 3 ( x ) is also an
eigenfunction of the translation operator and, if so, what eigenvalue of t R
pertains to it. In order to answer this question, we form tRfffpEklkzk3 =
t ~ p ~ k ~ k ~ k ~ (and a - 'obtain
~ ) the equation

= e x ~ l i P k , k ~ k - , ( ~ ~ ~ ~ R ) l ~ ~ ' ~ ~ k ~ k(2.117)
~ k ~ ( ~ ) .

The latter relation means that a p ~ k ~ k ~ isk indeed


~ ( ~ ) an eigenfunction of
the translation operator t R with eigenvalue e ~ p [ i P k ~ k ~ k ~ ( a -Evaluating
~R)].
,f3klkZ~3(a-1R) explicitly using equations (2.104) and (2.82), we have
90 Electronic structure of ideal C S ~ S ~ A I S

(2.118)
j i

This expression is subject to an interpretation which differs from the previous


one and which has important consequences for later developmentu. The new
interpretation is that the transformation, which hitherto operated on the
vector components of R, should now be understood to operate on the real
numbers kl,k z , k~ inslead. Of course, t,he new interpretation cannot change
the value of the expression (2.118). To assnre this, the kj must transform
according t,o the transposed matrix ';a of';a j n s t e d of the matrix a i j ,
which applies in the case of the components ~j of R. This is to say that

(2.119)

must hold. One may therefore write

(2.121)
3

This equation means that the transformation a,which originally operated


on the wavefunction p ~ k ~ l f a k ~ ( has
x ) , been transposd to operate 021 the
k,, which werp initially introduced as real numbers without any particular
transformation behavior. The transformation properties of the kj thusly
defined are similar, but not identical to, those of vector components. The
differenceliea in the fact that it ia not the matrix arj itself that multiplies
the column vector of the components, but rather it is the transposed inverse
matrix a;'. We will now prove that this is exactly the way the components
of a vector transform in a space reciprocal to the ordinary space of position
\-=tors x.

Definition of the reciprocal vector space

l h e space of position vectors x is defined through its basis al,a2,w. The


corresponding reciprocal vect.or space is determined by a basis bl, b2, b3 said
to be reciprocal to the original basis a1,az,a.The reciprocal basis is defined
by the set of equations

EQ ' bj = 2 ~ 5 i j (2.122)
These equations state that the vectors of the T & P T Q C ~ basis me normal to,
respectively, one of the three planes spanned by pairs of direct basis vectors.
2.3. General properties of stationary oneelectron states in a crystal 91

,4mong themselves, the bj are in general non-orthogonal to the extent that


the a, are mutually non-orthogonal. The lengths of the reciprocal basis
vectors bj are determined by the three equations which follow from (2.122)
for j = a. Altogether, the equation set (2.122) determines the reciprocal
basis vectors bj uniquely, aa follows:

2x 27r 2x
bl= -[a2 x q], bz - x a11 , b3 :-[a1 x az] . (2.123)
00 00 Qo

Here Ro = a1 . [a2x a31 is the volume of a primitive unit cell of the crystal
lattice.
If one s u bj et s the direct basis to a rotation or reflection, this induces
the same rotation or reflection of the rwiprocal basis, as follows immediately
from the defining equations (2.122) or (2.123). Because of these relations the
'reciprocal' basis (bl, bz, b3) is rigidy joined with the direct basis (al:a2,ag)
and will transform when the latter is tranaformd However, since the recip-
rocal basis is different from the direct one (only if (al,az,a3) are orthogonal
unit vectors are the two basis sets the same), the matrix LY which describes
the rotation or reflection of the a i is different from that which transforms the
bj. Using equation (2.123) one may easily show that the bj are transformed
with the transposed inverse of the transposed matrix u p which transforms
the q ,that is, with the inverse matrix a;'. In other words, the inverse
matrix describes the same rotation OF reflection of the reciprocal basis as the
transposed matrix docs fur the direct basis.
Consider now an arbitrary vector k of reciprocal vector space, which can
be written as

(2.124)
i

As we know, the vector components themselves transform in accordance with


the transpose of the transformation of the basis vectors. Since the transfor-
mation of the reciprocal basis calls for the inverse matrix, the components
bi will transform with the transposed inverse matrix 05'. This verifies our
earlier assertion that quation (2.121)describes the transformation of the
components of a reciprocal vector. Using this result, we obtain

kkb, = cxk. (2.125)


i

The reciprocal vector k of (2.124)may be used to express the phase ,&kka(R)


of the eigenvalue (2.103) of the translation operator in the more compact
form
92 Chapter 2. Electronic structure of ideal crystah

Bklkpl~3(R-
) -k. R. (2.126)
Using this expression the eigenvalue of t,hP treiistntion operator (2.103) be-
comes

which provides a more compact notation fm thp Bloch


We may write

(2.128)

) a short notatinii for the Bloch factor u ~ l ; ~ k ~ k ~ Because


with u ~ k ( x as (x).
of equation (2.115) one has

-
~ g k ( X R) UEk(X) . (2.129)
A further consideration illuminates the physical meaning of the reciprocal
vector k. For this p u r p o s ~we first assume the potential V(X) to be a constant
independent of x. appropriate to a free electron. All the abovementioned r e
s u l t s hold. of roursc, in this caw. Rut thrrp is m o m In this caw translations
through arbitrary vectors r and rotations about any axis and through arbi-
trary angles are symiwtry opcrations, for they do not change the potential
and, consequently, also leave the Hamilt onian innriant. Therefore.

"Ekb - 4= " E k W (2.130)


for arbitrary vectors r. This mcans that the Bloch factor ugk(x) is a con-
s t w t , and t,he eigenfunctious of the HamiItonian are of the form

(2.131)

This is just the well-known result that the stationary states of a reeelectron
may be taken as plane waves of a given wavwwtor k. If the potentid is not
completely constant but. as happens in a crystal, it remains constant only
under translations through lattice vectors, the meaning of k as wavevector
is largely prwervd. It is thrrefore called a qaasa-wavevector: Of course, the
dimension of the quasi-wsvevector is also that of a reciprocal length. The
Bloch functions y ~ k ( x )of (2.128) may be understood as travelling waves.
modulated spdiallJ by thr latticeperiodic Bloch factor U E ~ ( X ) .That the
stationary oneelectron states may be chosen in this form, is the content of
B l o c h ~thwrpm. This theorem d o e riot say, howevrr, that these states are
necessarily modulated tmvellzng plane waves, just as t h e stationary states
2.3. General properti= of stationary one-electron states in a crystd 93

Figure 2.4: Mesh of allowed


points in k-space due to the
periodic boundary conditions
in a 2-dimensional model.

of a free electron need not be travelling plane waves. One ran also h a m
standing plane waves, sphrriral waves d c .

Discretization of k-space

'The periodicity condition (2.96) for the eigenstates of the Hamiltonian has
largely been ignored so far. Fortunately, it can be satisfied very simply. The
Ebch fuucfions V E ~ ( X )obey the relation

if and only if the components k j of the k-vector are of the form


1
k,,- -
G
lj, j = 1!2 , 3 , (2.133)

with 4 BS arbitrary integers, to assure satisfaction of the requirement

The k-vectors defined by (2.133) form a finely meshed net in k-space (see
Figure 2.4). The only permissible k-vectors must be points of this net since
the Bloch functions are periodic with respect to the Periodicity region. The
k-space thus has a discrete structure; its set of points is countable. The
distance betwren different k-vectors is, however, so small because of the
large size of C that k can pract,ically be treat4 as it continuously varying
quantity despite its discrete character. This approximationwill be used later
extensively.

Reciprocal lattice

Just t t h in direct spaw, one may also consider a point lattice in reciprocal
space. In this context, the so-called mciprocal lattice i s especially useful. It
is dehed by taking the reciprocal basis vectors bl,bz,b3 of (2.123) aa its
94 Chapter 2. Electronic structure of ideal crystals

primitive lattice vectors. A reciprocal lattice vector K is then given by the


relation

K =CKjbj (2.135)
3
with arbitrary integers K j . As in coordinate space, one also has 14 different
lattice types in reciprocal space, namely the 14 Bravais lattices. The recip-
rocal lattices bear a close relation to the corresponding direct lattices. The
following properties can be verified easily:
(i) A reciprocal lattice has the same point symmetry as the direct lattice
with which it is associated.
(ii) The reciprocal lattice of a reciprocal lattice is the direct lattice.
(iii) If the direct lattice is primitive, then the same holds for the correspond-
ing reciprocal lattice. The reciprocal lattice of a centered direct lattice is
also centered.
The Bravais types of primitive direct lattices are the same as those of
their reciprocal lattices, while the Bravais types of centered direct lattices
and of their reciprocal lattices differ in certain cases, For example, a body
centered cubic reciprocal lattice corresponds to a face centered cubic direct
lattice, and, conversely, a face centered cubic reciprocal lattice corresponds
to a body centered cubic direct lattice.

2.3,4 Relation between energy eigenvalues and quasi-wave-


vector
We have not as yet directed attention to the question if there may be a
connection between the quasi-wavevector k and the energy eigenvalue E :
and what form it might take. In fact such B connection does e x i s t , and it
will now be explored in some detail.
That a Bloch function p ~ for k a given wavevector k cannot be art eigen-
function of I1 for an arbitrary energy eigenvalue E may be yeen as follows,
If one substitutes i p ~ k ( x )(2.128) into the Schriidinger equation (2.95),one
obtains the following eigenvalue equation for U E ~ ( X ) :

1
-(P
Izm
t rzk)2 -1- v(x)
1 UE&) = EUEk(X). (2.136)

The k-dependence of the operator in this equation transfers to the eigenvalue


E, i.e. E becomes H function E = E(k) of the quasi-wavevector k. In the
case of a free particle with V(x) = 0, one has E(k) = (H2/2m)k2. For
a crystal one expects, of course, more compljcated dependencies than this.
The specific form of the function E(k) is determined by the particular shape
of the periodic potential V(x). Some general properties of E(k) follow,
2.3. General properties of stationary one-electron states in a crystal 95

however, just from the space and time symmetry of V(x). We will discuss
this now, beginning with point symmetry, whose implications are relatively
easy to examine.

Point symmetry

We consider the Schrodinger equation (2.95), on the one hand, for the quasi-
wavevector crk, where TY i s taken to be an arbitrary element of the point
group of the crystal:

(2.137)

and, on the other hand, for the quasi-wavevector k, but with an application
of the transformation (Y as follows

HWEdX) E(kbPEk(X). (2.138)


The relations (2.120) and (2.124), derived in the preceding subsection, mean
that U p E k is an eigenfunction of H for quasi-wavevector ak. If there is
only orbe eigenfunction for a given wavevector, which we initially take to be
the case, a c p ~ k ( x must,
) be identical with ( P E ~ ~ ( x ),
because ( P E ~ ~ ( is
x )by
definition the eigenfunction for a k . Thus,

and a comparison of (2.137) and (2.138) then yields

E(k) = E ( a k ) . (2.140)
The assumption that there is only one eigenfunction of the Hamiltonian for
the quasi-wavevector k holds for almost all k. There is an exception for
those special k values for which ak does not differ from k or from a vector
+
k K equivalent to k for all a. One refers to such vectors as symmetrical k-
vectors. They will be considered in section 2.5, but we exclude symmetrical
k-vectors at this point. For non-symmetrical k-vectors, the fact that there
is only one eigenfunction of H , follows from the onedimensionality of the
irreducible representations of the translation group (see Appendix A).

Time reversal symmetry

A property of the Hamiltonian H which we have not employed thus far is


its invariance under time reversal. In order for the Schrodinger equation to
remain unchanged under time reversal, the time reversed wavefunction must
be defined as the complex conjugate of the original wavefunction. Then the
96 Chapter 2. Electronic structure of ideal crystals

time reversed Bloch function of wavevector k has wavevector -k. Since both
Bloch functions have the same eigenvalue, one obtains

E ( k ) - E(-k). (2.141)

This relation has additional significance, beyond that of relation (2.140)


obtained by means of spatial symmetry, only then, if the point group of the
crystal does not contain the inversion.

Translation symmetry: Extended and reduced zone schemes

The lattice translation invariance of H has a remarkable consequence for


the eigenvalue function E(k). It arises in conjunction with the degeneracy
of the eigenvalues ck(R) of the translation operator tR. Specifically, all
wavefunctions (PEk+K with K , any reciprocal lattice vector, have the same
eigenvalue of thc translation operator. Independently of the actual value of
K, the latter is
-2k.R (2.142)
Ck+K(R) = e

which follows directly from the relation

K .R- (2.143)
1

with K , and T % as integers. Therefore we obtain the full spwtrum of eigen-


values of t R if k varies within a particular primitive unit cell of k-space. The
corresponding k-vectors will br denoted by kl. The k-vectors of any other
primitive unit cell differ from these by a reciprocal lattice vector K and thus
do not lead to new eigenvalues of the translation operator. However, the en-
ergy cigenvalues g ( k ) do differ, so that E(k1) f h(k1-t- K ) holds in general.
Table 2.3 illustrates these connections.
This asymmetry with respect to the k-dependence of rk(R) and E(k) is
the reason for the introduction of two different representation schemes for
the energy eigenvalue function E(k). In the representation scheme employed
thus far, k varies over the entire reciprocal space, and to each vector k we
associate only one k-dependent energy function E(k). With respect to the
eigenvalues of the translation operator, this description i s highly redundant,
since one encounters the same eigenvaliie an infinite number of times. This
disadvantage is eliminated if the selection of the region over which k varies
is not fitted to the energy E ( k ) but rather to the ck(R) eigenvalues of the
translation operator, i.e., if k is allowed to vary only over a primitive unit cell
rather than over the whole k-space. Then we must accept, however, that an
infinite number of different energy eigenvalues are assigned to each k-vector,
namely all those which follow frotn E ( k ) by means of the prescription
2.3. Gcnmd propertic3 of stationary oneelectron states in a crystal 97

Table 2.3: Representation of band structure in the cxtcnded and reduced zonc
schemes.

I
Wavevector Eigenvalue of tR Eigenvalue of H Represent at ion

I
[-ik. R]
~ X PE(k + K) General
~ exp [-ikl . R]

k from
1Extended
Infinite Zone Scheme
Space

k = kl from
Unit
' Reduced
Zone Scheme
Cell

whme kl is a vector of the primitive unit cell at K 0. In this way the


~

singlevalued function E(k) over all k-space is transformed to an infinitely


multi-valued function EK(k1) over a primitive unit cell. An analogous state-
ment holds for the eigenfunctions of the Hamiltonian. Each Bloch function
p,yk over all k-space corresponds to an infinite number of different Bloch
functions pEKkl(x) - pqkl+Iqkl+I{(x). One cells the representation in-
volving the entire k-space the extended zone scheme, and that involving the
primitive unit call at K = 0 t h e w d u d zonw schpme. For most problems
in semiconductor physics the reduced scheme is more convenient than the
extended scheme.
Later, we will demonstrate that the energy E(k), as a function over all
k-space, is not continuous everywhere but has discontinuities at certain k-
planes. Anticipating this, one may then inquire whether one can choose the
primitive unit cell of the reduced scheme at K 0 in such a way that the
~

discontinuities of the branches EK(k) do not occur in the interior, but only
on the surface of the primitive unit cell. If such a special primitive unit
cell exists at all, it must have at least the symmetry of the point group of
direclionb, as followb from the symmetry reletion (2.140). This nieans that
only the Wigner-Seitz cell of the reciprocal lattice is a possible candidate for
siich a primitive unit cell. Of course, its point group is that of the crystal
lattice and, theIefore, larger than the requisite point group of directions.
98 Chapter 2. Electronic structure of ideal crystals

We know, however, that lattices have fewer point groups than there are
point groups of equivalent crystal directions. The next-lower lattice point
group would not contain the needed point group of directions. The issue
in question thus revolves upon whether the function branches EK(k) are
continuous in the interior of the Wigner-Seitz cell of the reciprocal lattice at
K = 0. We will prove that this is indeed the case for the branch Eo(kl) in
the next section. Moreover, it will be shown that the other branches EK(k1)
of the infinitely multi-valued function E(k) may be redefined in terms of
new branches such that each single-valued branch is also continuous over
the Wigner-Seitz cell at K = 0. One calls these continuous branches energy
bands and the Wigner-Seitz cell of the reciprocal lattice at K = 0 is the f i r s t
Brillouin zone.
Having started from general considerations and obvious conjectures, we
thus arrive at the important conclusion that the spectrum of allowed energy
values of crystal electrons will have the form of energy bands, and that these
energy bands arise by varying the quasi-wavevector k over the first Brillouin
zone. We will discuss the concept of Brillouin zones along with the still
outstanding proof of the continuity of E(k) over these zones more fully in
the next section.

2.4 Schr6dinger equation solution in the nearly-


free-electron approximation
It is well known that the Schrodinger equation can be solved approximately
if the potential, or a part of it, represents a small effect which may be treated
by means of perturbation theory. To apply this procedure, the solution of
the unperturbed problem must be known. At this point, we will treat the
entire periodic potential by means of perturbation theory. Even if it is not
to be expected that this will yield results which are quantitatively accurate,
some qualitative understanding can be gained in this way.
If one takes the periodic potential as a perturbation, then the unper-
turbed problem is that of a completely free electron. The corresponding
Hamiltonian Ho is simply the kinetic energy of the electron, i.e.

H o = -.P2 (2.145)
2m
The perturbing operator H I , which together with Ho forms the total Hamil-
tonian
H = Ho Hi, + (2.146)
is then the periodic potential

H I = V(x). (2.147)
2.4. Schriidinger equation solution in the nearly-freeelectron approximation 99

In perturbation theory, the exact eigenfunctions Cpk and eigenvalues E ( k )


are expanded in power series,

+
E(k) = Eo(k) 6E1(k) + 6E2(k)+ . . . (2.149)

with respect to the perturbation potential, here with respect to the periodic
potential V(x). In this discussion, we suppress the energy index E of the
eigenfunctions for the sake of brevity. This leads to no ambiguity since we
work in the extended zone scheme, where the energy is a unique function of
the quasi-wavevectors k. The zero-th order terms obey the equation

Hop; = Eo(k)Cpg. (2.150)

In order that the perturbation theory be applicable, the zero-th order solu-
tions must form a complete orthonormal set of functions, such that

and
-yCpp(x)Cp;(x) = 6(x - x) (2.152)
k
must hold. The k-summation of (2.152) is extended over all points of the
infinite finely meshed net of Figure 2.4. The two relations (2.152) and (2.153)
are in fact valid for the solutions

(2.153)

of equation (2.150). The validity of the orthonormality relation (2.151) may


be easily verified by direct calculation, and the completeness relation (2.152)
is proved in the theory of Fourier series. The energy eigenvalue Eo(k) of
zero-th order for pg is
0 Ti2
E (k) = -k2 (2.154)
2m
We will assume at the outset that this eigenvalue is not degenerate, more
exactly, that this holds for k-values for which the perturbation operator V
has non-vanishing matrix elements (pi! I V I ~ p g )with any vector k. It is
not possible to exclude degeneracy completely because wavevectors k of the
same length lead to the same energy eigenvalue (2.154).
100 Chapter 2. Electronic structurc of ideal crystals

2.4.1 Non-degenerate perturbation theory


h accordance with the procedures of quantum mechanical perturbation t h e
ory, we have

(2.155)

(2.156)

The energy correction 6E1(k) of the first order i s the average value (cpg I
V I pi) of the periodic potential. This value may be set equal to zcro. The
off-diagonal matrix element of the potential involved in (2.1551, (2.156) are
given by

(z.157)

Because of the lattice periodicity of V(x) we can transform the integral over
the periodicity region in (2.157) into a sum o f integrals over the primitive
unit cell Qo(0) at R = 0:

Noling that
(2.159)

(2.160)

we arrive at the result

Using this result, the correction 6 ~ :to the wavefunction takes the form

(2.162)

and the total wavvcfunclion p k = p! t 6pk is given in this approximation as


2.4. SchrMingw equation solution in the nearly-fieet3-electronapproximation 101

One recognizes that this has the expected Bloch form (2.128) b r energy
eigenfunctions of the crystal Hamiltonian. Moreover, (2.163) yields the Bloch
factor as

(2.16 4)

The second order energy correction E2(k) may be determined using (2.161)
as

(2.165)

The corrections to the eigenfunctions (2.162) and the enerffi eigmvtrlues


(2.165) will become large if the denommator in these expressions becomes
srnall. Actually. the denominator may even vanish when

E"(k)- Eo(k I K ) - 0. (2 166)


In such a case, I ~ basic
P premise for validity of non-degenerate perturbation
theory is violated, namely that there be no degeneracy among the unper-
turbed energy cigenvalues for those k-values fur which the matrix elements
of V ( x ) do not vanish. However, just such a degeneracy exists if equation
(2.166) holds, with the states 9: and cp:-K degenerate with each other.
Below, we will accommodate this case and discuss the corresponding degen-
erate perturbation theory, but first we analyze the k-values for which the
condition (2.166) is fulfilled. Using (2.154) for E o ( k ) ,we get from (2.151)
1
k.K+ -IK
2
12= 0. (2.16 7)
This equation can be solved geometrically. It is obvious from Figure 2.5
that the tIps of the corresponding k-vectors lie on a plane perpendicular
to the vectors K or -K, with this plane intersecting the vector -K at its
half-length. Such planes are familiar from the theory of X-ray diffraction by
crystal lattices. They aTe called Bmgg reflectam planes.
The non-degenerate perturbation-theoretic results discussed above may
be interpreted quite clearly if one regards the problem of the calculation of
the eigenfunctions in the periodic potential as a scattering problem. In this,
the unperturbed eigenfunctions p: are thought of as incoming waves and
the perturbed eigenfunctions 9 k as outgoing waves, 6 ~ corresponding
: to
the scattered part of the latter. That the perturbation theory according to
formula (2.162). yields only minor corrections to the unperturbed eigenfunc-
tions is related in this picture to the fact that the incoming and scattered
+
plane waves in\-olve different wavelengths, since 1 k If jk KI, and cor-
respondingly they are not capable of significant constructive interference.
102 Chapter 2. Electronic structure of ideal crystals

-&

\ -K
f

Figure 2.5: Construction of Bragg Figure 2.6: Illustration of perturba


reflection planes. tion theory with respect to the peri-
odic crystal potential in the vicinity of
a Bragg reflection plans where the zero-
th order energy levels are degenerate.

For incoming waves with k-vectors close to the Bragg reflection plane of the
reciprocal vector -K, the scattered part dpk contains a p a r t i d a r l y large
plane wave component of wavevector k + K. This is due t o the fact that,
in this case, the incoming plane wave and the Ecaltered wave of wavevec-
tor k + K, have almost the same wavelengths. They are thus rapable of
constructive interference which strongly enhances the amplitude of this par-
ticular scattered wave. If k is located exactly on the Bragg reflection plane,
one may understand the scattered wave with wavevector k +
K as the plane
wave reflected at this plane in the sense of geometrical optics. This follows
immediately from the construction of the k h g g reflection plancs in Figure
2.5. The scattered wave of wavevector k +K is strengthened by inter-
ference, w h c h makes it the refkected wave (bear in mind that in the wave
picture reflection represents an interference phenomenon). This interpreta-
tion can be universally applied to the propagtltion of plane waves of any kind
in a system of scattering centers periodically ordered on a lattice, including
X-rays. Actually, the above results for elwtron waves were discoverd iisinE
X-ray radiation.
The amplitudes of plane waves reflected at Bragg planes, according to
formula (2.162), become idnitely large formally What is expectd. how-
ever. is that the amplitudes of the refiected and incoming waves should be
comparable. This contradiction arises from the misuse of non-degenerate
perturbation theov, w h h is no longer applicable under the conditions of
reflection. For k-vectors which obey the Bragg condition (2.167), one must
apply degenerate perturbation theory.
2.4. Schrdinger equation solution in tho nearly-free-electronapproximation 103

2.4.2 Degenerate perturbation theory

Let ko be a point on a Bragg reflection plane, so that


1
ko.K+ - IK 12= 0 (2.168)
2
holds for a particular reciprocal lattice vector K. We denote a small deviation
from ko by Ak where 'small' means that I A k I<<] K I. A vector k of the
form
k=ko+Ak (2.169)
is then a vector close the Bragg plane involved. The same holds for the
+
vector k K, as it lies close to the parallel plane in Figure 2.6.
According to quantum mechanical perturbation theory for the case of
degenerate zero-th order states, one proceeds as follows. The perturbed
eigenfunctions cpk are sought as linear combinations of the two (almost)
degenerate eigenfunctions cpi
and (p& in zero-th order, writing

(PIFo(x)= cocp:(x) +C K d + K ( X ) (2.170)

with co and CK coefficients to be determined. In the lowest non-vanishing


order of perturbation theory, the Schriidinger equation leads to the following
set of equations;

Eo(k) - E
( V(K)
V(K)*
Eo(k$ K ) - E
(2.171)

For the set to have a non-trivial solution, its determinant must vanish,
whence

Eo(k) - E V(K) = 0
(2.172)
V(K)* Eo(k+ K ) - E
must hold. The two solutions are E = Ek(k) with

1 1
Ek(k) -- ,[Eo(k)
2
+ Eo(k + K)] f -\/[Eo(k)
2
- Eo(k + K)]*+ 4 I V(K) I*.
(2.173)
For further evaluation of this expression, we assume that Ak is directed
parallel to K (see Figure 2 . 6 ) . Expanding the square root in powers of
I Ak I and terminating the series with the second order, we find

jpK2
E*(k) = Eo(ko)f I V(K) I 4 li2 [
1 f 2m
I V(K) I] Ak2.
(2.174)
104 Chapter 2. Electronic structure of ideal crystals

Figure 2.7: The degeneracy of the two energy parabolas E(k + Ak) and E(k +
K + Ak) at the Bragg reflection plane Ak = 0 (left part of figure) is removed by
the periodic crystal potential. In the formerly continuous energy spectrum a gap
arises at this plane (right part of the figure).

The remarkable feature of this result is that the 2-fold degenerate energy
+
eigenvalue Eo(ko) = Eo(ko K ) at the Bragg reflection plane, i.e. for
Ak = 0, splits into two different eigenvalues which are shifted by IV(K)I
to higher and lower energies, respectively. Between them there is a region
of width 2 I V ( k ) I where no allowed energy values may exist (see Figure
2.7). This means that, in the formerly continuous energy spectrum of the
free electron. an energy gap appears as a result of the periodic potential.
This occurs at all Bragg reflection planes, provided that the corresponding
Fourier components I'(K) of the periodic potential do not vanish. Except
for the Bragg planes, the function E ( k ) and the energy spectrum remain
continuous. The result derived here by means of perturbation theory is
also valid well beyond the framework of validity of this approximation. The
following statement remains true also in the general case.
The energy f u n c h o n E(k) LS continuous everywhere in k-space, with the
exceptzon of the Bragg reflectLon planes where, 271 generat, discontinuities
occur and the energy spectrum exhihts gaps.

In regard to the shape of the function E*(k) close to the Bragg planes,
the relation (2.174) provides the following picture (see Figure 2.7). The two
branches E+(k) and E-(k) approach Ak = 0 with horizontal tangents to
+
the limiting values Eo(k) IV(K)I and Eo(k)- ll'(K)I, the upper branch
with positive curvature, and the lower with a negative one. The latter ob-
servation follonrs from (2.174) since the applicability of perturbation theory
is restricted in validity by 1 V(k) I << h 2 K 2 / 2 m ,which means that the sign
2.5. Band structure 105

of the angular bracket in (2.174) is always determined by the second term.

2.5 Band structure


The perturbation-theory calculations of the preceding section were carried
out in the extended zone scheme for E(k}. Alternatively, there is also the
reduced scheme introduced in section 2.3, in which the quasi-wavevector k
varies o d y over a primitive unit cell of reciprocal space. In this, each of an
infinite number of choices for this cell can be used. Here, we will determine
the choice of primitive unit cell of the reciprocal lattice in such a way that the
desired property of the function E(k} discussed above, namely that it have
discontinuities only at Bragg reflection planes and otherwise be continuous,
is described as simply as possible.
Simplicity of the description calls for the discontinuities of E ( k ) to occur
only on the boundaries of the unit cell, with E(k)continuous in the interior.
The question of whether there is a primitive unit cell which guarantees that.
is equivalent to the question of whether there exists a primitive unit cell
which is bounded by Bragg planes, but has no other such planes in its inte-
rior. The prescription for the construction of the Bragg planes (see section
2.4) makes it immediately clear that the Wigner-Seitz cell of the reciprocal
lattice at K = 0 has the desired property - it is bounded completely by
planes which obey equation (2.1671, namely Bragg planes, and in its interior
it is devoid of such planes. This means that within the Wiper-Seitz cell of
the reciprocal lattice at K = 0, the function E(k) is continuous. However, it
remains to be clarified how the rest of k-space can be reduced t o the Wigner-
Seitz cell at K = 0 in such a way that the resulting new function branches
for E(k) are also continuous over this celL It is clear that this cannot be
done by simply dividing k-space into Wiper-Seitz cells and then translating
all cells not containing the origin through reciprocal lattice vectors back to
the cell at K = 0. The reason for this is that the non-central Wigner-Seitz
cells are cut by Bragg planes. The correct procedure goes back to Brillouin
and will be described below.

2.5.1 Brillouin zones

Definition

Consider, in k-space, centro-symmetric bodies which contain the origin and


are entirely bounded by parts of Bragg reflection planes. These bodies are
arranged and enumerated according to their volume. The smallest will be the
Wigner-Seitz cell at K = 0. The second, next-largest body, has the volume
of two Wigner-Seitz cells, as one may easily demonstrate. It contains the
106 Chapter 2. Electronic structure of ideal crystals

ZONE 1 2 3 4 5 6 7 8 9 1 0

Figme 2.8: The first, 10 Rrillouin zones or H sqimre plane 1al.tic:e. (After Brillouin,
1953).

first body jointly with the Bragg planes bonnding i t . If one removes the
first body from the second, then one obtains a hollow body of the volume of
one Wigner-Seitz cell, which is bounded by Bragg planes inside and outside,
but contains no siirh planes in i t s interior. This conhtruction may be carried
+
further. Removing the v-th body from the (v 1)-th one again obtains a
hollow body having the volume of the primitive unit cell in k-space. It i s
boiindd by Bragg planes, and has no such planes in its interior. These
difference bodies are called Brdloum zones (abbreviated 8 8 s ) . The v-th
body i s called the v-th R%, the Wigner-Seitz cell at K = 0 is, accordingly,
the first B Z . In Figure 2.8 the first 10 Brillouin zones are shown for a
2-dimensional square reciprocal lattice. The first B Z for two important 3-
dimensional lattices, i.e. the frr lattice and the hexagonal lattice, are shown
in Figure 2.9. We will derive the shape of the first of them below.
Focusing on a point k, of the v-th R Z , a v e t o r K(k,) of the reciprocal
lattice is attached k, such that k,- K(k,) represents a vector kl of the first
BZ,

kl = k, - K(k,). (2.175)
2.5. Band structure 107

Figure 2.9: The first Brillouin zones of the two Bravais lattices which apply to many
semiconductors: (left) the face centered cubic lattice of diamond, zincblende and
rocksalt structures; (right) the hexagonal lattice of wurtzite and selenium structures.

We omit the formal proof for this important result here. For the square
lattice it follows immediately from Figures 2.8 and 2.10. Its validity in the
general 3-dimensional case is clearly plausible. If, instead of k,, another
point kh of the v-th B Z is chosen, another reciprocal lattice vector K(kh)
may be necessary in order that kh - K ( k l ) shall be a vector of the first
B Z . The set of reciprocal lattice vectors K(k,), K(kL), . . . which arises if
k, varies over the whole v-th B Z , is just the set of K-vectors which defines
the internal boundary planes of the v-th B Z . The correspondence of the
points of the v-th B Z to points of the first B Z given by equation (2.175)
is unambiguous in both directions - each k, corresponds exactly to one kl,
and each kl exactly to one k,. One terms this correspondence folding of
the v-th B Z into the first. The term displacing would probably provide a
better description. During such folding or displacing, the external or internal
boundary planes of v-th B Z may be translated to the interior of the first
B Z . There, they border on other boundary planes of the v-th B Z . This
means that original kl-vectors which lie immediately to the right and to the
left of such planes will involve different reciprocal lattice vectors.
The folding operation (2.175) provides the mechanism for changing from
the extended description of the energy function E ( k ) over all infinite k-space
to the reduced description within the first B Z . The transfer of the function
values is, defined above by relation (2.144). It states that the value E(k) at
a point k = k, of the v-th B Z is assigned to that particular point of the
first B Z which arises from k, by folding the v-th B Z into the first. By this
assignment a multi-valued function E,(kl) is defined within the first BZ.
The unique function E(k) over the whole k-space is mapped into the multi-
valued function E,(kl) within the first B Z . Formally, this is expressed by
108 Chapter 2. Electronic structure of ideal crystals

Figure 2.10: Folding of the second through ninth Rrillouiri zones into the first R Z
for the plane square lattice of Figure 2.9. (After Brillouin, 1953)).

the equation

Ev(ki)- E ( k i + K(k,)) E(kv). (2.176)


We maintain that the branches E,(kl) of the multi-valued function are con-
tinuous within their entire range of ddinition, i.e. within the fust B Z . For
points kl which correspond to points k, of the interior of the v t h B Z ,
the validity of this assertion is obvious. For those psrticiilar kl-planes sris-
ing from a bo~iirlaryplane of the v-th B Z , continuity is not immediately
evident, for in penetrating such a kl-plane the original vector k,, of the Y-
th R Z jumps by the negative of the rcciprocal latticc vector. which defines
this boundary plane. However, this jump does not affect the energy eigen-
value E(k,) in the extended zone scheme as may be seen by means of the
degenerate perturbation tbwry in swtion 2.4. Thus the functions E,(k) are
also continuous on the particular planes in question.
Each of the function branches F;,(kl) defined above encompasses some
finite interval of values on the energy scale. The term energy hnnd for the
set of E,(kl)-values is thus obvious; I / is the so-called band andm Between
the various function branches, or energy bands, energy regions may exist in
which no energy eigenvalues occur. These regions are called forbidden zone8
or energy gaps. l'he whole s r t of functions R,(kl) is referred to collectively
as thr band structure.
In Figure 2.11 we illustrate the band structurc using the model of a
1-dimensional lattice. The solid curves correspond l o a rornplelely free elec-
tron, and the dashed ones to an elwtzon in a weak periodic potential. One
recognizes that energy gaps orcur not only on the boundaries of the first BZ
2.5. Band striictiire 109

Figure 2.11: Rcduction of the energy parabola of a free electron into the first B Z
of a 1-dimensional lattice. The changes caused by a weak, periodic lattice potential
are indicated by dashed lines.

but also at its center at k = 0. This is due to the fact that, during folding,
surface points of the first B Z are displaced to the B Z center, and the energy
gaps from those points also appear at the center.
The folding procedure into the first B Z must also be carried out for the
eigenfunctions of the Harniltonian. Let p v k l denote the Bloch function of
energy E,(kl) and of reduced wavevector kl. In the extended scheme the
same wavefunction reads q q k , ) k v ( x ) . Thus,

The representation of E(k) as a multi-valued function E,(kl) over the first


B Z is by far the most common and frequently used description of the energy
spectrum of a crystal. The first Brillouin zone is, therefore, of outstanding
importance in solid state physics. Unless otherwise specified, the quasi-
wavevector will henceforth always be understood to lie in the first B Z . Thus
we will write k instead of kl, which means, in particular, E,(k) instead of
E,(kl). The Schrodinger equation for an electron in a crystal then takes the
form
110 Chapter 2. Electronic structure of ideal crystals

f f ~ v / c ( x=
) &(k)pvk(x). (2.178)
The first RZ is the fully symmetric primitive unit cell of the reciprocal
lattice. The relation between the reciprocal and the direct lattices was dis-
cussed in section 2.3, where it was seen that each of the 14 direct Bravais
lattices corresponds to a particular reciprocal lattice of the same symmetry,
although not necessarily of the same Rravais type. Accordingly, there are
14 different reciprocal Bravais lattices, which ah0 means 14 different first
B Z s . Two of them, the two that are most important for semiconductors,
are shown in Figure 2.9.

Empty-lattice band s t r u c t u r e

In Figure 2.11 the energy E ( k ) of a free electron moving in 1 dimension


has been represented over the first B Z of R I-dimensional lattice. Similar
representations are possible in 3 dimensions. Below we demonstrate this for
the face centered cubic lattice which applies to crystals of the diamond and
zincblende structures, According to Table 1.2, the primitive vectors of this
lattice may be chosen in the form
a a a
2 2 2
.
a1 = -(O, 1,I ) , a2 = - ( l , O , I), a 3 = -(I, I, 0) (2.179)

The corresponding basis vectors of the reciprocal lattice follow from equation
(2.123), with the result
27r 27r - 2n - -
bl = -(I, T, I), b2 = -(l, 1,T), b3 = -(I,
a 1,l). (2.180)

The first BZ is bounded by the 14 planes perpendicular to the following


reciprocal lattice vectors:
(1) 8 lattice vectors having the length of the primitive lattice vectors given
above, which means the smallest possible length overall. These vectors read:

2n 27r - 2x - - 2n
-(i,i, T), -(I, 1,T), -(I, 1,I), -(T,T,T),
a a a a
27r - 2s 2n 27r
-(1, 1,1), -(I, T, I), -(I, 1,T), -(I, 1,i). (2.181)

(2) 6 lattice vectors having the next-smallest possible length:

27r 27r 2n
-4%
a 0, O ) , -(0,2,0),
a -(O,
a 0,Z). (2.182)
2.5. Band structure 111

Figure 2.12: Symmetry points and irreducible parts of the first BZ of the fcc
lattice (left) and the hexagonal lattice (right).

The first B Z of the fcc lattice obtained by means of these reciprocal lattice
vectors is shown in Figure 2.9. Perpendicular to the three cubic axes it is
bounded by squares, and normal to the four space diagonals by hexagons. It
is common to denote the symmetry points of the first BZ by capital letters.
Greek letters are assigned to symmetry points in the interior of the first BZ,
and latin letters to symmetry points on its surface (see Figure 2.12). The
center of the B Z , for example, is I?, the point at which a cubic axis cuts the
B Z surface is X , and the connecting line between I? and X is A.
We will now reduce the energy parabola E(k) = (fi2/2m)k2 of a free
electron along the A-line, i.e. for points

r = 2R
-(o,
a
o,o) 2R
A = -(o,
a
2?r
o , ~ ) ; x = -(o,
a
0, 1) (2.183)

with 0 < C < 1. In the reduced zone scheme we have

E,(k) = -h2
(k
2m +K ( ~ Y ) ) ~ . (2.184)

Using the above listed 14 reciprocal lattice vectors K(k,) in this relation,
the second and third BZ's are folded into the first. For convenience we
~ an energy unit, and set
introduce Eo = ( T ~ ' / 2 m ) ( 2 7 r / a )as

E,(k) 71 E u . ~v(k). (2.185)

With lattice constant a of 5 A, a value of 5.9 eV follows for Eo. For ty(k)
we have
112 Chapter 2. Electronic structure of ideal crystals

(2.186)

where the K,, Kvy,K,, are the components of the vectors K, in units of
2nla. In Table 2.4 we show the reduced energy functions ty(<) for the 14
vectors K, together with those for 0. In some cases the same value of the
function E,(C) is obtained for 4 different K,s, meaning that the correspond-
ing energy bands are 4-fold degenerate. In Figure 2.13 the content of Table
2.4 is illustrated graphically, along with the band structure parallel to the
A-line between 0 and L not given in Table 2.4. Again, one recognizes that
there are k-points associated with the same energy value several times.
The band structure shown in Figure 2.13 is that of an empty fcc lattice,
i.e. a fcc lattice whose lattice points are not occupied by atoms, since the
latter would create a non-zero periodic potential V(x)if they were present.
For a vanishing potential V(x),no definite lattice exists because of its arbi-
trary translation symmetry. Nevertheless, a definite lattice, the fcc lattice,
and a definite lattice constant, were chosen in the above procedure. This
was an arbitrary choice, in the sense that we could have also chosen any
other of the 14 Bravais lattices, and any other lattice constants. However, if
we want the empty lattice band structure to resemble the band structure of
a really existing crystal, then we cannot take any empty lattice band struc-
ture but must chose the one for the lattice applying to the crystal under
consideration. Later we will verify that the band structures of real crys-
tals and the pertinent empty lattice band structures have in fact common
features. In Figure 2.11 such features are, for example, the appearance of
the lowest energy eigenvalue at the I?-point and the development of several
non-degenerate energy bands from this level along the A-line, one of them
crossing another band at the X-point. It is also consistent that the energet-
ically high-lying bands display a relatively high degeneracy at the I?-point,
although for real fcc crystals the degeneracy at maximum can be only 3-fold,
and not 8-fold as for the empty fcc lattice at the energy E = 3&, and not
&fold as for the fcc lattice at E = 4Eo. A degeneracy higher than %fold
turns out be accidental, i.e. not caused by symmetry, but by the particular
values of the potential, which is identically zero in this case. The width of
the lowest energy band in Figure 2.11 is about Eo, i.e. 6 eV. Also, this
result is not too far from the actual widths in fcc crystals, as we will verify
below.
As with the reduction along A- and A-lines, one can plot the band struc-
ture along other symmetry lines in the interior and on the boundary of
the first B Z . A complete representation of the functions E,(k) over the
points of 3-dimensional k-space would require a Cdimensional space. In the
3-dimensional space available to us it is only possible to represent bands
E,(k) over planes in k-space. Such representations are, however, uncom-
2.5. Band structure 113

(KV&V,K,) cy(c) Notation Degree of Degeneracy

( i i i ) (iii) (111) (iii) 2 + (6 + I)* c2 4

(zoo) (200) (020) (020) 4 I cz F3 4

mon. -4s a rulc, one plots the energy against lines in k-space,as in Figure
2.13.
The question of what valws the energy band functions E,(k) take out-
side the first B Z does not arise or is meaningless, since the E,(k) were

"L A r A
Wavevector - X

Figure 2.13: Band structure of an empty fcc lattice.


114 Chapter 2. Electronic structure of ideal crystals

just defined by transferring the function values from other B Z s to the first
one. For this reason, the E,(k) already encompass the entire spectrum of
eigenvalues as k varies over the first B Z . If one wishes to also consider the
functions E,(k) outside of the first B Z , one has to define them there. The
most natural manner to do this is the periodic continuation.

Periodic continuation

In this context, the values of E,(k) in Wigner-Seitz cells having centers at


K # 0 are defined by the relation

(2.187)

where k is a point of the first B Z . By periodic continuation, one can of


course extend each function originally defined only within the first B Z to the
entire k-space. The question is what analytic properties does this function
have? If one considers an arbitrary function, discontinuities will occur on the
boundaries of the Wigner-Seitz cells, i.e. the periodically continued function
will not be continuous in the entire k-space. In order to have continuity,
E,(k) must satisfy certain conditions on the boundary of the first B Z . These
conditions follow from the fact that the boundary of the first B Z consists
of pairs of equivalent parallel planes (see Figure 2.6). We denote a point on
one plane of such a pair by ko, and kb denotes the equivalent point on the
other plane of the pair. If ko belongs to the first BZ,as we will assume, then
kb cannot also belong to it, because a primitive unit cell contains only non-
equivalent points. The value of E,(kb) is defined by the periodic continuation
and, as such, it is equal to E,(ko). In order for the periodically continued
func%iorrt o be continuous, thix value must coincide with the limiting value
of E,(k) if one approaches the point kb coming from the interior of the first
B Z . Thus, a necessary condition for continuity of the periodically continued
function is that the original function, defined only over the first B Z , obey
the boundary condition

lim E,(k) = E,(ko), (2.188)


k- k&

where k is in the interior of the first B8. This condition is also siifFicient,
since the continuity of the periodically continued function E,(k) implies
that equation (2.188) holds. Therefore, the boundary condition (2.188) for
the function E,(k) defined over the first R E , and the continuity of the
periodically continued function E,(k) are just different descriptions of the
same property. Tising continuity of the function &(k) defined over all k-
space, this property may be expressed in a more convenient way. This is the
reason why one continues the e n e r a bend fiinctions R,(k), which actually
2.5. nand structure 115

have meaning only within the first BZ,periodically over the entire k-space.
One could also forego this, and use the condition (2.188) directly.
We will now prove that the energy bands E,(k) actually obey the bound-
ary condition (2.188), and thus, are continuous everywhere as periodically
continued functions. To this end, we use the Schrodinger equation (2.178)
for the eigenvalues E,(k). The corresponding Bloch functions p,k(x) may be
expanded in plane waves with wavevectors k + K , where K is an arbitrary
reciprocal lattice vector, as follows:

(2.189)

Using this result, the Schrodinger equation (2.178) takes the form

(2.190)

The E,(k) are the eigenvalues of the matrix of coefficients of this system of
equations and their k-dependence i s determined by that of the matrix. The
latter is manifpstly continiioiis. That it also exhibits the periodicity of the
reciprocal lattice, one may verify as follows. Consider the coefficient matrix
+
of (2.190) at the point k K with K an arbitrary reciprocal lattice vector.
If one replaces the column and row indices Kand K , respcctively, by K+K
+
and K K, using the relation (K KIVJK+ + K) - (KIVIK), one
obtains the matrix at the original k-vertor. This implies that the eigenvalues
Ev(k t K) and E,(k) are identical, as stated.
The above proof employs only the lattice translation symmetry of the
potential V(X). The continuity of the periodically rontinued function E,(k)
in the entire reciprocal space is therefore an immediate consequence of the
periodicity of V(x) in the direct lattice. The continuity of the periodically
repeated energy band functions E,(k) is often very useful. It ensures, on the
one hand, that the E,(k) may be npproximctted fairly closely by a Fourier
series with relatively few terms. On the other hand, mathematical theorems
dealing with the types and numbers of extrema of periodic analytic functions
may also be applied. This i s particularly important for the theory of the
optical proprrties of semiconductors.
We return now to a question which was previously explored in section
2.3, namely the implications of point symmetry for band structure. This was
discussed above in the extended zone scheme, and symmetric k-points were
excluded. NOW we USP the reduced zone scheme description and concentrate
on symmetric k-vectors.
116 Chapter 2. Electronic structure of ideal crystals

2.5.2 Degeneracy of energy bands


The first B Z , as the Wiper-Seitx cell of the reciprocal lattice, also cxhibits
the point synmelry of this lattice which, for its part, is the same as the point
bymmetry of the corresponding direct lattice. The latter statement follows
directly from the definition (2.123) of the reciprocal basis bl, bp, b3. The
first B Z of the fcc lattice has, therefore, the point symmetry of a cube, and
its point group is Oh(m3m). The energy bands likewise have a particular
point symmetry, namely that of the equivalent directions of the crystal. This
is expressed by quation (2.140), which underlies the extended zone scheme.
It may, however, be transferred immediately to the reduced scheme, so that
we may also write

&(k) = E v ( a k ) , (2.19 1)
where a is an element of the point group of equivalent directions of the
crystal. This group is generally smaller than the point symmetry group of
the lattice, but in special circumstances it can also be the same. The latter
rase occurs for the diamond structure, where the point group of equivalent
crystal directions is likewise o h .
In section 2.3, symmetric k-vectors, i.e. vectors which are transformed
by at least one element o f the point group 1 of equivalent directions, into
themselves or into an equivalent vector, were excluded. Now we also admit
such k-vectors to consideration. In r q a r d to their effect on symmetric k-
vectors, the elements cr of P split into two subsets. The Erst contains those
elements which transform k into itself or, at points on the surface of the
first B Z , inlo a vector crk equivalent to k which differs from k only by a
reciprocal lattice vector K. This set forms a subgroup pk of P , and is called
the s n ~ a l lpoznt group of k. The second subset contains all those elements
of P which transform k neither into itself nor into a vector equivalent to k.
The set of all different and non-equivalent vectors a k is called the star ofk.
We denote the number of elements of the point group P by p , and the number
of different star points of k by S k . Since each star point has the same point
symmetry, the number pk of elements of the small point gToup Pk is given by
the relation p k p l s k . A general k-vector bas no point symmetry, whence
7

p k = 1 and A k = p . For k-vectors on symmetry lines or planes, p k has a


value between 1 and p . The center of the first BZ is special: all elements of
P arc also dements of Zk, so that pk = p and S k = 1. A wavevector whose
small point group does not consist of only the unity element, is a symmetry
point. The points of the first B Z in Figure 2.12 marked with Iargr greek or
latin letters are among them.
Using the terminology above, the point symmetry of band structure as
given by equation (2.191) may also be described by saying that the energy of
a particular band has the same value at all points of the star of a wavevector,
2.5. Band structure 117

Onc therefore also calls this symmetry the Y~UT degunrrucy OJ e n e r g y bands.
Bmause of star degeneracy, the energy band functions E,(k) need to be
calculatcd only for a scction of the first B Z which covers the region between
adjacent star points (Figure 2.12). One calls such a section an zrrediiczble
part of the first R E . The energy eigmvulues for the remainder of the first
B Z are obtained by means of the symmetric continuation (2.191) of the
values of the irreducible part. The irreducible part is the p-th part of the
first BZ. If one considers that, in the case of diamond structure, the number
p of elements of the point group 01, is 48, one can appreciate how greatly the
description of band structure is simplified by exploiting the point symmetry
of crystals.
In section 2.3 it was shown that for any eigenfunction (p,k(x), acp,k(x) =
cpVk(cr-x) is also an eigenfunction of the Haniiltonian H with the same
eigcnvalue B,(k), provided (1 belongs to the symmetry group of 11. Those
elements a of this group which also do not change the wavevector k, or
change it only by a reciprocal lattice vector, will thus transform an eigen-
fiinction pvk(x) of a particular Land v and k-vector, into an eigenfunction
cp,k(~u-~x) of the same band and the same k-vector. Thus, lor the case in
which not all wp,k(x) are linearly dependent, t h e are several linearly in-
dependent eigenfunctions for a given band index 11 and wavevector k. Let
their number be d,k. This means that t h e energy band E,(k) i s d,k-fold de-
generate at the point k. One calls this kind of degeneracy band degeneracy.
We employ q l k ( x ) ,1 - 1 , 2 , . . ,duk, to denote a basis set in the subspace
of eigenfunctions or H having the eigenvalue E,(k). The functions crcplk(x)
are then likewise eigenfunctions with the eigenvalue E,(k). As such, they can
be expressed as linear combint\tions of the basis functions pVk(x),whence

(2.192)

with Dpl(a) as expansion coefficients. Through relation (2.192) each element


a of the sniall point group is associated with t~ matrix D p l ( a ) . In Appendix
A we explain that the matrices D p l ( a ) form a r e p r e s e n t a t z o n of the small
point group of the wavevector k. If the clegeneiacy of the energy eigenvalues
i s only due to the symmetry of the Hamiltonian, then this representation
is arreducihle. We may say that the eigcnfimctions of the Hamiltonian for
a particular k-vector and band-index v span a subspace in Hilbert space
which gives rise to an irreducible representation of the small point group.
The dimension of this reprcsentation, i.r. the dimension of its matrices,
corresponds to the degree of degeneracy of the energy band for the k-vrctor
mider consideration. If one knows all irreducible representations of the small
point group Pk, then one also knows what degrees of degeneracy of the energy
bands are possible at k. This relation between degeneracy and symmetry
118 Chapter 2. Electronic structure of ideal crystals

Table 2.5: Irreducible representations (notations and dmensions) of the smaLl


point groups of symmetry points and lines of the fist B Z of the fcc lattice for
crystals with the diamond structure. For the point X, the projective irreducible
representations are given in the crystallographic factor system. (See Append& A ) .

-
A
-
X

-
c
K

is quite remarkable. It holds not only for the energy bands of crystals,
but for the eigenvalues of any Hsmiltonisn in quantum mechanics. It is a
good example how a relatively abstract mathematical theory - the theory of
groups has immediate physical consequences.
~

The number of distinct irreducible representations of an arbitrary finite


group is, b i t e , as are the dimensions of the representations. For the point
groups of equivalent directions, the irreducible representations can be only
1-,2- and 3-dimensional; l-dimensional for small groups, 2-dimensional only
for groups that are 'not too small', and 3-dimcnsional only for the largest
point groups, specifically for Oh(m3m),0(43m), Th and Td(332) of the cubic
crystal system.
To illustrate these somewhat abstract statements and to prepare for the
discussion of actual band structures in section 2.8, we consider some s p a i d
k-vectors of the first BZ of the fcc lattice. At the center r, the small point
group is identical with the full point group P for crystals of diamond
structure, i.e. with o h . In the context of energy band theory, the irre-
ducible representations of the small point groups are denuted by the same
capital greek or latin letters which stand for the k-vectors, supplemented by
subscripts. For the I?-point and diamond structure, the irreducible repre-
sentations are denoted by rl,Tz,I112,r25. r15, and I?;, I?$, I&, This
somewhat strange indexing refers back to the so-call& compatibility rela-
2.5. Band structure 119

tions between different representations (see Appendix A for more detail).


The primes indicate that the involved representations differ from the un-
primed ones only by a minus sign for the inversion matrix. In Table 2.5, all
irreducible representations of the point group o h , which is the small point
group of r for crystals with the diamond structure, are listed. In addition
to the symbols of the representations, their dimensions are given. The irre-
ducible representations of the small point groups Pk of the symmetry points
A, X , A and L for crystals having the diamond structure are also indicated
in Table 2.5. According to this table, one has both 1-, 2- and %fold degen-
erate bands at the center of the first BZ,while at all other points only 1-
and 2-fold degeneracies occur for crystals which have diamond structure. At
X there are 2-fold degenerate bands exclusively. This peculiarity is due to
the fact that the space group of the diamond structure is non-symmorphic.
In this case, the irreducible representations of the small point group of X
involve a factor system, which means that multiplication of two representa-
tion matrices yields the matrix of the product element, save only for a scalar
factor (see Appendix A). In such circumstances, it is understandable that
1-dimensional representations may not be possible.
The degeneracy at I' may be compared with that of a free atom. In the
latter case the degeneracy is likewise a consequence of rotation and reflec-
tion symmetries. However, arbitrary rotations and reflections are possible in
this case, i.e. the point group contains an infinite number of elements. The
dimensions of the irreducible representations of this group are determined
by the angular momentum quantum number I , where 1 can take the values
+
0,1,2,. . . ,00. Amounting to 21 1, all odd numbers are possible as dimen-
sions of irreducible representations, and hence also as the multiplicities of
energy eigenvalues. In the case of the hydrogen atom, the symmetry-related
degeneracy is still not the full degeneracy. In addition, one has an acciden-
tal degeneracy caused by the particular shape of the Coulomb potential. In
this case the energy eigenvalues differ only for different principal quantum
numbers n. Thus all eigenstates corresponding to a given n , with their vari-
ous 1-values 0, 1 , 2 , .. . ,n - 1, have the same energy. The degeneracy is thus
n2-fold for the hydrogen atom.

2.5.3 Critical points and effective masses


The degeneracy of energy bands at symmetry points of the first BZ which we
treated in the preceding subsection, is a direct consequence of the symmetry
of these points. We proceed now to another consequence of this symmetry,
which will lead us to the concept of critical points and effective masses.
If a particular symmetry point ko of the first BZ is transformed into
itself under the action of a point symmetry operation a , then points ko 6k +
in the vicinity of ko are necessarily transformed into points close to ko. If,
120 Chapter 2. Electronic structure of ideal crystals

for example, a i s a reflection with respect to u plane normal to the x-axis,


+
then points of the form ko 6 k e, with ex as unit vector in x-direction, will
transform into ko - 6 k . ex. Because of the symmetry relation (2.191) for
energy bands, E,(k) will have the same value before and afler the reflection.
From this it follows that the first derivative of the function EV(k) with
respect to k, must vanish at the mirror plane. If ko also simultazlcously lies
on a second mirror plane perpendicular to the first, e.g., one perpendicular to
the y-axis, thrrr the b,-derivativp of f!!,(k) will also be zero. If ko i s even more
symmetric and involves a third mirror plane normal to the z-axis, then the
derivative with respcrt to k, also has to vanish. What is being demonstrated
here using the example of mirror planes has a general significance: some or all
of the first derivatives of the energy band functions E,(k) with respect to the
three wavevwtor components vanish at symmetry points, depending on the
degree of symmetry. With some ambiguity of expression, we may say that at
symmetry centers all three fiist clerivativcs are zero, at symmetry lines two
vanish, namely those with respect to the normal plane, and on symmetry
planes one derivative vanishes, nainrly the one in the normal direction. On
a symmetry line there are also often special points, wheie, for reasons that
have nothing to do with symmetry, the remaining first derivative parallel to
thv line also vanishes.
The first derivatives of the function Ey(k) with respect to k have a direct
physical meaning. One can easily prove (and we will do so in section 3.3)
that they are, apart from a constant factor, equal to the expectation value
of the momentum operator p in a Bloch Btate, thus

(2.193)

If VkE,(k) vanishes, then the average momentum in the corresponding Bloc11


state is also zero.
These are occasions which warrant special consideration and examination
of points the first B Z in which all first derivatives of E,(k) vanish simulta-
neously. These points are called crztical points. We will later see that the
so-called denszty of + d a t w possesses singilnrities at energies whose constant
energy surfaces contain critical points. This also explains the name of these
points. The critical points h a w a close connection with symmetry points,
without, however, being generally identical to them. Critical points k, are
defined formally by the equation

(2.194)

With the requirement that E,(k) be a regular function at k,one may


approximate it in the neighborhood of k, by a Taylor expansion to second
2.5. Band structure 121

order

Here, we introduce the components M;:g of a second rank tensor M L 1


defined by the equation

(2.196)

IJsing this tensor, the expansion (2.195) of E,(k) may be written in the form

(2.197)

Thc tensor components cnter this expression in a manner similar to the


way the reciprocal mass rn-l enters the energy dispersion relation of a free
electron. Actually one can obtain the expression (2.154) for the energy of a
free electron from (2.197) by Erst taking the energy zero at E u ( k c ) which is
unimportant, and then by substituting

(2.198)

The components of the invrrsr tensor M , of M; accordingly have the di-


mension of a mass. One therefore calls M u the effectzve mas8 tensor. The
term effective refers to the fact that the electrons of a crystal are not free
elcctrons but arc subject to the influencc of thc periodic potential. T h e
presence of this potential mandates that the wavevector of the free electron
be repIared by the quasi-wavevwtot k, and that the quadratic dppendenre
of the energy on k become a more complicated dependence. Only in the
vicinity of critical points does the quadratic drprndenre persist, but with
coefficients M;$, which differ from those of the flee electton. They are
eflectztw coefficients which involvc the effrct of the periodical potmtial.
In contrast to the free electron case, where the coeficients form a multiple
rn-l of the unit tensor 6,p, i.e. a scalar quantity, for the case of an electron
in a crystal they represent u tensor ML$ with non-vanishing off-diagonal
elements and general diagonal elements. An analogous statement holds for
the inverse, as the cffective mass tensor M,. Thus, in general, the effective
mass is a tensorial quantily. Therein the crystal potential manifests itself,
generally generating different mass values in different spatial directions. In
special cases one ran approximate the effertive mass by a scalar quantity
rn;. Then, the difference between the mass of the free electron and that of
the crystal electron reduces to the different magnitudes of the two masses.
z 22 Chapter 2. Electronic structure of id& crystals

The mathematical description of effective mass can be simpl%d by exploit-


ing the symmetry of the tensor Mv:, with rcsppct to an exchange of its
indices a and 3, i.e. by using the relation

(2.199)
This propat,y follows directly from the defining equation (2.196) for M&.
As is well-known from linear algebra, such a real symmetric tensor cau be
brought l o diagonal form by a coordinate transformation, in our case of the
components of the k-vector, to the principal axis system. With ms' being
the principal diagonal elements of M l l in this system, one has

The m;;' are real. Ifn general, they can take both positive and negative
values. The same applies to thc inverse quantities, the effective masses rnk
themselves. Taking k, as componenls of k in the principal axis system, we
have

(2.201)

If all three effective masses are positive, the energy function E w ( k ) has a
minimum at the critical point kc,and if all three are negative it is a maxi-
mum. If the three effective masses do not all have the same sign, then E , ( k )
has a saddle point.
Consider the particular case in which the critical point lies on a 4-
fold symmetry axis of the first 3 2 . and that this is simultaneously one of
the principal axes. Then, for symmetry reasons, the two principal tensor
elements normal to this axis must be equal. With LY = 3 for the symmetrical
principal axis we thus have

mC1 = m t 2 (2.202)

The corresponding element for the symmetrical principal axis m:3 E m:,,in
general differs from mZI. If k,, however, lies on a symmetry center of the
first B Z , then all three elements are in fact identical, i.e.

(2 203)
For the frequentIy observed case of a symmetry point at the center of the
first BZ!i.e. with k, = 0, one gets the simple dispersion law

E,(k) = E v( 0) + -k2.
F12
(2.204)
2m;
2.5. Band structure 123

This differs from that of a free electron only in that the effective mass rn;
appears instead of the free mass and that the energy at k = 0 assumes a
value E,(O) which depends on the band index and, therefore, cannot be set
equal to zero for all bands by changing the zero of energy.
In the following subsection we will again deal with band structure in its
general form E,(k). We will introduce the so called density of states that
encompasses essential information on band structure in just one function of
energy. Often, a knowledge of the density of states is enough to calculate
important macroscopic properties of semiconductor crystals.

2.5.4 Density of states


Definition

We first consider an arbitrary infinite oneelectron system. Its Hamiltonian


will be denoted by H , its eigenstates (disregarding spin) by li), and its energy
eigenvalues by Ei, i = 1 , 2 , .. . , 00. Periodicity of the eigenstates is assumed
with respect to a periodicity region of volume 51. The quantity

2
p ( E ) = - C 6 ( E - Ei). (2.205)
O i
is called the density of states (DOS) of the system. We will show that this
designation is justified, i.e., that p ( E ) does represent the number of o n e
particle states of energy E per unit energy and volume. To this end we
integrate equation (2.205) over a small energy interval E < E' < E A E , +
obtaining

P(E)AE
=
2
FL E+AE
dE'6(E' - E i ) . (2.206)

The integral on the right hand side yields

1 for E < E i < E + A E


JEE+AE dE'6(E' - Ei) =
0 otherwise .
(2.207)

+
Thus, each state i whose energy lies between E and E A E contributes '1'
to the sum on i, with no contributions from all other states. The sum in
(2.206) equals, therefore, the number A z ( E )of states having energy in this
interval. Considering spin, one has

2
~ ( E ) A=E- A z ( E ) , (2.208)
51
124 Chapter 2. Electronic structure of idcal crystals

confirming the designation of p ( E ) as DOS.


The L)OS expression (2.205) may br written in a more compact form
which will be used in Chapter 3 in the derivation of important general results.
Noting that the 6-function 6 ( & ) is the imaginary part of -(1/7r)l/(E +
ie)
with t a small positive number, one can easily prove that p ( E ) of (2.205),
except for a factor R, equals

p ( E ) = --Im Tr (2.209)
7T

where trace sytxho1 Tr denotes the sum over all diagonal elements with
respect to any basis set in Hihert space. If this set is identified with thr
eigmstatcs 12) of ff, expression (2.209) transforms immediately into (2.205).
We now consider the DOS p ( E ) defined by (2.205) for an ideal crystal.
The cigenstates li) are Rloch states Ivk), in this case, and the DOS p ( F )
therefore reads
2
-
p(E)-
C 6 ( E-
uk
(2.210)

The k-sum is over all points of the finely meshed net of Figure 2.4 which lie
within the first B Z . Because of the fineness of this network, the result of the
summation is almost identical to that of an integration. The substitution of
the s m by an integral is done in accordance with the prescription

k
-1
c...= 8T3
R d3k ...
(2.211)

The factor f l / 8 m 3 in front of the integral ocrurs because the volume of a


mesh element is the G3-th part of the volume 8n3/Slo of a primitive unit cell
of the reciproral lattice, whirh is 87r3//n. One must divide by this volume in
writing the sum as an integral. The density of states follows as

(2.212)

which is independent of the volume R of the system. The density of states


restricted to the u-th energy band, p u ( E ) ,is defined by the expression

(2.213)

and it differs from zero only for energy values for which the argument of the
6-function may brrome zero, i.r. for energies in the v-th band. Summing
over d l bands one again obtains the entire density of states

(2.214)
2.5. Rand structure 125

Expression (2.213) for p , ( E ) may be transformed into a surface integral over


the constant energy surface &(k) -= E using the rules of calculation for the
&function, thus obtaining

(2.215)

with df as element of the constant energy surface E,(k) -= E . The inteErand


l/IVkF,,(k)l is, apart from a factor TL, the reciprocal absolute value of the
group velocity of an electron in the Bloch state Ivkq). The smaller this
velocity is at a particular k-point on the constant energy surface or, in other
words, the longer an electron stays at it, the larger is its contribution to the
density of states. If (VkE,(k)l - 0, i.e. at a critical point, the contribution
is, formally, infinitely large. From this it follows that for energy values E
whose iso-energy surfaces contain critic81 points, singularities of the density
of states as a function of energy are to be expected. They are called van
Hove singularitiee.

Parabolic energy bands

We aasunie an isotropic parabolic energy band, i.e. we suppose that the


function E,(k) is given by the expression (2.201) with the sainc effective
mass rn; along all three main axes. The critical point k,is taken as the
origin 0. If we further replace E,(k,) by E N , then the expression for &(k)
becomes

(2.2 16)

We use this to calculate the density of states p , ( E ) from equation (2.213).


The effective mass rn; is taken to be positive, i.e., E,(k) should have a
minimum at k = 0. The use of the dispersion law (2.216) i s only justified
within its limits of validity, i.e. for energy values E,(k) close to the minimum
E d . Only for these E-values can the density of states as calculated by means
of (2.216), be relied upon. For such energies it so happens that the value of
p , ( E ) does not depend on the upper limit of integration. We transfer this
limit, for simplicity, from the edge of lhe first B Z to infinity, obtaining

(2.2 17)

In this integral, we transform to new variables x,y, z


To start we, execute the integration over z using the rule
126 Chapter 2. Electronic structure of ideal crystals

1 [S(z - ZO)
E - EA - x 2 - y2 - z 2 ) = - +6(z+ ZO)] (2.2
2zo
with

Zo = ,/E - E , , ~- x 2 - y2 f o r E - E& > 0. (2.2


It follows that

where

1 f o r E > Euo
B(E - Euo) = (2.221)
0 for E < Euo
is the Heaviside unit step function. If one considers, instead of an isotropic
parabolic band, an anisotropic one with three different effective masses
m:l, m:2, m:3 corresponding to the principal axes of the effective mass ten-
sor, then the density of states is again given by expression (2.220), but m:
must be replaced as follows

(2.222)

by the so-called density-of-states-mass m b . This may be seen immediately


if one reviews the calculation for the isotropic case. The anisotropic effec-
tive mass involves a change only upon substitution of the variables for the
components of the k-vector - the factors for the 3 components are different
and lead to the modification indicated in equation (2.222).
According to expression (2.220) the density of states of a parabolic band
with m*, > 0 exhibits square root-like behavior at the lower band edge, and
continues following a square root law up to infinitely large energies. The
latter property is a consequence of our untenable assumption of parabolic
dispersion up to the upper band edge and the replacement of this edge by
the point at infinity. In reality the band edge lies at finite energies, and the
density of states again falls off to 0 there. It has, qualitatively, the shape
shown by the dashed curve of Figure 2.14. The sudden square root-like
increase of the DOS at the band edge reflects the fact that the function
p(E) has singularities at these energies - the first derivative with respect to
E is 0 if one approaches the edge from the forbidden zone, and it is +CXJ if
one approaches it from the interior of the band. This is one of the already
mentioned van Hove singularities of the DOS. These singularities occur not
only at energies where a particular band E,(k) has a minimum, as in the
2.5. Band structure 127

Figure 2.14: DOS of a parabolic 5- I I I I

isotropic band in the vicinity of its min-


imum (solid curve). The dashed curve
shows the DOS of a more realistic band.
-c 4-

= 3-
<
I

vl
2-
0
0
1- -

0
-1 0 1 2 3 4
Energy lorb. unlts)

case considered here, but at all energies corresponding to critical points of


the energy bands, thus also at maxima and saddle points.

Counting states: 3D, 2D,and 1D density of s t a t e s

The square root enerw dependence of the density of states in the case of a
parabolic dispersion law may also be demonstrated in the following, more
vivid way (also see Figure 2.15). For simplicity we set E,o = 0. and omit
factors which do not depend on k or E. The n m b r r A Z of b m d states
with energy between E and E .f A E is the number of different k-points of
the finely meshed net which yield energy values in t,his interval. These k-
values lie in a thin spherical shell in k-space [see Figure 2.15) which, because
E o(. k2, has radius k~ 6: &.and thickness Ak which, since Ah' CK k&k,
is given by

1
Ak 0; -AE. (2.223)
fi
For the volume AV of this shell, it follows that

AV m k L A k cc A A E . (2.224)

Since the density of the finely meshed net of k-points of the first BZ i s the
same everywhere, the number of k-points in the spherical shell is propor-
tional to its volume AV. Hence AZ cy v%AE and

p(E) o(. 4%. (22 2 5)


Such considerations can easily be applied to energy bands in 2-dimensiouel
(2D)and 1-dimensional (1D)k-spaces. Such k-Bpaces have physical (as
well as mathematical) meaning, in particular for electron systems whose free
128 Chapter 2. Electronic structure of ideal crystals

b C

Figure 2.15: Counting states in (a) 3D, (b) 2D, and (c) 1D k-space.

motion is confined to one or two dimensions of 3-dimensional space. We wiU


encounter examples of such systems in Chapter 3 - surfaces and quantum
wells for the 2D-case, and electrons in a homogeneous magnetic fieId for the
1D-case. For 2 0 k-space, the volume of the spherical shell is replaced by
the area A F of a circular ring, such that A F a k E A k LX A E and

p ( E ) 3: const. (2.226)
The density of states of a 2D-electron system is therefore independent of E.
In the case of a ID k-space: AV is replaced by the length Ak of the k-interval
itself. Thus,
1
p ( E ) DC -AE. (2.227)
fi
2.5.5 Spin
Thus far, spin has been omitted from our discussion of the general properties
of stationary oneelectron states of crystals. It turns out. however, that spin
and spin-orbit interaction play an important role in most semiconductors.
Therefore, the question arises as to how the results derived above for scalar
wavefunctions change when electron states are no longer scalar but spinor
functions and the spin-orbit interaction H , is taken into account.
The starting point to address this question is the one-electron Schrodinger
equation (2.58) with spin, which may be written in the following short form

(2.228)
2.5. Band structure 129

The first point we will consider is that of spatial symmetry in the presence
of spin.

Spatial symmetry

Inspecting the explicit form of the spin-orbit interaction operator ( b e y for-


mula ( 2 . 5 6 ) ) , one readily recognizes that TKso has thc full symrwtry of the
crystal. Thus it colrimutes with all elements 9 of the space group. Since the
same holds for the spin free part H of the total Hamiltonian !i+ one
has

Iff + H,. g] = 0 . (2.229)

To exploit this symmetry property of thv total Hamiltonian w e mu61 know


how thc components p(x, s) of a spinor transforin i d e r thr operations g of
the space group. Considering translations first. we have

In t.he absence of spin, the cornmulivity of the HanliltoIlittn with translations


gave rise to the Bloch theorrm. In t.he same way that this theorem was
proved without spin above. its validity may be also demonstratd here - in
4).
the pwsrnrr of spin. It. stdates that the solutions { p ~ k ( x , y ~ ( xf)}
, of
t.hr Schrodingeer equation (2.228) for the eigenvalue E A ( ~can ) be chosen in
the form of Bloch type spinor functions with a particiilar quasi-wavevecbor
k as

zti)
where I ~ X ~ [ X . are the spin-dependent Bloch factors. If k is restricted to
the first B Z . then E x ( k ) rrprwents a continuous function of k. Thc band
index X here refers to both the state of the orbital motion and also to spin
state.
Second, we consider point symmetry operations u , whose action in tram-
forming t,hc components of a spinor is discussed in Appendix -4. The results
of Appendix A are applied here in the form
130 Chapter 2. Electronic structure of ideal crystals

with

In this, $, B and p are the Euler angles of the orthogonal transformation a.


If the spinor is spatially constant, the transformation just reduces to multi-
plication by the matrix D r ( a ) . The set of all orthogonal transforniations cy.
2
forins an (infinite) group. Through relation (2.233) this group is assigned a
matrix group D i ( a ) . A peculiarity of thm mapping is that it is not unique
1
- a change of the angles cp or $J in (2.233) by 27r leads to a change of the
sign of the associated matrix, even though this signifies the identity trans-
formation. One may say that under a rotation through 27r a spinor does not
transform into itself, but into its negative. This gives rise to multiplication
rules for the representation matrices which deviate from those of the group
itself. Thus the representations of the full orthogonal group in the space of
two-component spinors are not representations in the ordinary sense, but
in a generalized sense. They form projective representatiom with a special
factor system (see Appendix A), or, in short, spinor representatiom. By
means of the so-called double group, which includes each element, of the fill1
orthogonal group twice, once in the original form, and once multiplied with
a rotation through 27r, one may trace back the spinor representations to or-
dinary representations. The spinor representations are those representations
of the double group which do not occur among the ordinary representations
of the single g ~ o u p(for the derivation of this result see Appendix A).
Up to this point we have only given attention to the transformations of
spinors in spin space. The states of electrons are described by spinor fields
having positional dependence. These give rise to spinor representations of
the (full) orthogonal group which differ, in general, from D i , The total-
ity of spinor representations may be obtained from the ordinary irreducible
representations V, and the particular spinor representation D1 of the full
2
orthogonal group. Indeed, if a spinor field (P,(x, s), with s = fk,transforms
according to a certain ordinary representation D, in coordinete space, then
its transformation in coordinate spin space is governed by the product
~

representation Vi x Vv.If, here, D, encompasses all vcctor representations,


2
then all spinor representations are obtained.
Using the concept of spinor representations, the already discussed con-
nection between the eigenfunctions of the crystal Hamiltonian for a given
eigenvalue and the irreducible representations of its symmetry group may
be generalized in the following way:
The spinor eigenfunctions of the crystal Hamiltonian W +
H , having the
same energy eigenvalue span a representatdon space of a n irreducible spinor
2.5. Band structure 13 1

Table 2.6: Irreducible spinar representations (notations and dimensions) of the


small point groups of symmetry points and lines of the firat B Z of the fix lattice
for crystala with the diamond structure.

representation of the full symmetry group of H + H,, (namely, the space


group of the crystal).
The product representations Dr.x V v , taken as representations of the sman
point groups of a particular wavevector rather than for the full orthogonal
group, are generally reducible. This means that bands which are degener-
ate without spin by reason of spatial symmetry, may split if spin is taken
into account. The magnitude of the energy splitting depends on the spin-
orbit interaction, without this interaction, degeneracy persists, but as an
accidental rather than a symmetry induced degeneracy. The following ex-
ample illustrates the removal of degeneracy. We consider the valence band
of zincblende type semiconductors which is triply degenerate at I'. The
product representation in this case is Di x 1-15. It decomposes into the
a
two irreducible spinor representations ra and r 7 of the point group T d of I
',
where rs is of dimension 4 and r7 of dimensiun 2. This means that, due
to the spin-orbit interaction, the triply degenerate r15-valence band without
spin splits into the $-fold degenerate rs-band and the 2-fold degenerate r7-
band, accounting for the effects of spin. In Table 2.6 the irreducible spinor
representations are shown for some symmetry paints of the first BZ of dia-
mond type semiconductors. A systematic treatment of these representations
is given in Appendix A.
132 Chapter 2. Electronic structure of ideal crystals

Time reversal symmetry

As the consequences of spatial symmetry are altered by the phenomenology


of spin, so is the action of time reversal symmetry changed by spin. First,we
will show that the operator K which-transforms a spinor {cpx(x, $), cp~(x, $)}
into the spinor K{cpx(x, i), cpx(x,i)} at reversed time, is given by the rela-
tion

(2.234)

To prove this assertion, one executes the time reversal operation on the
spin-orbit interaction operator H,, of equation (2.56) explicitly. The vector
a' then transforms into -a*, and p into -p, so that the net change in Hso is
the replacement of d by a'*. Employing the operator K , on the other hand,
the effect of timc reversal on H , , may be expressed in terms of the similarity
transformed operator KH,,K-'. For the two expressions to be identical,
K must have the form given above in equation (2.234), apart from a phase -
factor which remains undetermined. The wavefunction {cpx(x,
- i), cpx(x,$)}
must be replaced by K{cpx(x, i), cpx(x, $)} for the Schrodinger equation
(2.180) to remain valid.
The question, under what conditions time reversal symmetry includes
degeneracy of eigenfunctions not accounted for by spatial symmetry alone,
is treated in Appendix A in full generality. Here we deal only with a specjal
case. We consider a non-symmetric wavevector k. Let {cpxk(x, i),cp~(x, i)}
be an eigenfunction with energy Ex(k). If the point group of directions -
contains the inversion, then the eigenfunction {cpx-k(x, i), cpx-k(x, i)}
cor-
responds to the same eigenvalue Ex(k) -= Ex(-k). The time reverse of
the Bloch function {cpx-k(x, i), cpx-k(x, $)} likewise has energy Ex(k) and
-
i),
wavevector k. It is linearly independent of {(Px~(x, cpxk(x, $)}, since

as one may easily show by evaluating the scalar product (here, one also has to
sum over the spin coordinate s ) . Thus, two linearly independent eigenstates
of the same energy have the wavevector k. Since k was chosen arbitrarily, it
follows that, due to time reversal symmetry, all bands of inversion-symmetric
crystals are at least twofold degenerate.
2.5. Band structure 133

2.5.6 Calculational methods for band structure determina-


tion

There are many methods, quasi-analytic and numerical, for calculating the
band structure of crystals. In the following we will give an overview.
In band structure calculations one is confronted with two problems,
firstly with the formulation of the one-elcctron Schrodingcr equation, i.c.
with the determination of the effective periodic oneelectron potential V ( x )
of the crystal, and secondly with finding the eigenvalues and eigenfunctions
of that equation. The various methods of band structure calculation differ
in the manner in which thcsc two problcms are resolved. Hcre we will deal
mainly with the second problem, i.e. with the solution of the Schrodinger
equation whose potential is known. The first problem, the determination
of thc cffectivc oncclcctron potential V ( x ) , has in principle brcn solved
in section 2.2, where we dealt with the one-electron approximation for the
many-electron system of a crystal. Here, we only address some further dctails
of this problem.

Determination of the effective one-electron potential

The simplcst way of dealing with thc oneelectron potential V ( x ) is to treat


its matrix elements with respect to a particular basis set as empirical param-
eters rather than integrals to be evaluated from a knowledge of the particular
profile of V ( x ) . Having done that, one has an empzmcal methodof band struc-
ture calculation. We will provide examples below. If one wants to forego,
however, the aid of empirical data and calculate the band structure from first
pnnczples, this evasion is unacceptable. The effective one-electron potential
V ( x ) of the crystal must then be known as a function of x. Methods which
proceed in this way are referred to as ab znztzo methods.
Within the frozen core approximation, V(x) describes three interactions
of a valence electron. The f i s t is the Coulomb interaction with the atomic
nuclei - this part is not problematic. The second is interaction with the
(frozen) core electrons - the energy levels and eigenfunctions of the core
states of the free atoms have to be determined in advance for us to be able
to write down this potential contribution explicitly. It includes the Hartree
and exchange potentials as well as the correlation potential of the valence
electron-core electron interaction. These potentials are described by ex-
pressions which are completely analogous to those for the valence electron-
valence electron interaction derived in section 2.2. The third part of the
periodic crystal potential accounts for the effect of the remaining valence
electrons on the valence electron specifically considered. It depends on the
very same eigenfunctions that are to be calculated. Because of this, the
calculation of the band structure must be done self-consistently. In many
134 Chapter 2. Electronic structure of ideal crystaIs

cases, the Hartree and HartreeFock approximations fail to give satisfying


results, so that correlation effects must also be inchided. As indicated in
section 2.2. one way this can be done is by means of the dematy finctional
theory in its local approximation (local density approximation or LDA). The
LDA-method yields very good results as far as valence bands are concerned,
but it fails if applied to the conduction bands. In particular, as already
mentioned in section 2.2, it does not reproduce the correct d u e of the fun-
damental energy gap. A procedure which avoids this failure is the Gmens
functzon method mentioned in section 2.2. This approach is now being ap-
plied more frequently under the name puma-particle method (see, e.g., Bech-
stedt, 1992). Within this framework the self-eneqy operator is often taken
in the so-called GW -approximation [G stands for Greens function and
W for the Coulomb potential).
We add two further remarks, related to the potential of the atomic cores.
Firstly, only for relatively light atoms such as C or Si, may spin and the
spin-orbit interaction of the valence electrons be ignored. For the heavier
atoms, such as Ge, this interaction is essential and must be incorporated
in the effective one-electron potential. Secondly, the decomposition of the
total electron system into valence and core electrons need not to be made in
the literal sense of these terms. What counts is which electrons are frozen
in their atomic states and, therefore, need not be treated self-consistently,
and which electrons must be. The latter are valence electrons in a more
general sense. In 111-V semiconductors, for example, they can also include
d-electrons which, of course, do not belong to the valence shell of one of
the two elements involved. In the extreme case, all electrons are treated as
valence electrons. Then one has the so-called call electron problem. The
solution of this problem involves an extraordinary large numerical effort, so
that such all-electron band structure calculations have been performed in
o d y a few cases to date. From the physical point of view, they are the most
satisfying. With increasing computing power they will become ever more
important.

Solution of the Schrodinger equation

The solution procedures can be divided into three groups, firstly! the matrix
methods, secondly, the cell methods, and thirdly, the muffin-tin methods.

Matrix methods

In the application of matrix methods one represents Bloch type eigenfunc-


tions of a given quasi-wavevector k in a particular basis set consisting of a
finite number of functions of the Bloch type. The Hamiltonian of the crystal
is thereby represented by a k-dependent complex Hermitian matrix which
2.5. Rand structure 135

has as many columns and rows as the basis set has functions per k-vector.
The calculation of band structure is thus traced back to the determination
of the eigenvalues and eigenvectors of a k-dependent Hamiltonian matrix.
The basis sets employed differ by the number and kinds of functions
they contain. One would like to manage with the fewest possible functions,
to minimize the numerical effort. The price for this is a loss of precision,
because with fewer basis vectors the eigenfunctions are necessarily approxi-
mated more crudely than with a larger number - in the extreme, one needs
an i n h i t e number of functions. For a fixed number of basis functions the
level of precision achieved is higher for basis functions which are better ad-
justed to represent the eigenstates. The following basis functions prove to
be of practical use:
-Plane waves Iki-K) with k being a vector of the first BZ and K a reciprocal
lattice vector. This constitutes a generalization of the nearly free electron
approximation.
~ Bloch sums of atomic orbitals, also called LCAO's (Lznear Combznatzons
of Atomzc Orbatuls), or of other localized functions. This includes the tight
binding met hod.
~ Bloch functions with Bloch factors for a special wavevector, referred to as
Luttznger-Kohn funetzons. The so-called k . p-method uses these functions.
In some circumstances, the tight binding and k pmethods may be used
to derive analytic expressions for the k-dispersion of energy bands. Such
expressions are extremely useful to achieve a physical understanding of band
structure. Therefore, we treat these two methods in greater detail below
(sections 2.6 and 2.7).
- Orthogonalized plane waves, called OY W's. An OPW-function IOPW~+K)
is obtained from n plane wave Ik+K) by subtracting a certain linear combi-
nation of core band eigenfunctions Irk) of the crystal Hamiltonian. Within
the frozen core approximation, the core band eigenfunctions (ck) may be
taken as Bloch sums of the core states of single atoms. The linear combi-
nations to be subtracted are chosen such that the OPW's are orthogonal to
all core eigenstates, so that

(2.236)
holds. The OPW's must have the form

+
l O l ' W k + ~ )= Ik. K ) - C(cklk + K ) Ick) (2.237)
C

in order to satisfy the orthogonality condition (2.236). Making the expan-


sion functions orthogonal to the core eigenstates accounts for the fact that
136 Chapter 2. Electronic structure of ideal crystals

eigenstates of the Hamiltonian for different energies are always mutually


orthogonal. This means that the sought-after eigenstates of the valence
electrons of the crystal for a given quasi-wavevector k must be orthogonal
to the core eigenstates having the same quasi-wavevector k. For this reason
the O P W ' s are much better adjusted to represent the eigenfunctions of the
valence electrons than are pure plane waves; correspondingly one needs fewer
O P W ' s than plane waves to accurately represent the valence eigenfunctions.
A further development of the OPW-method is the pseudopotential method.

Pseudopotential method

The idea underlying the pseudopotential method is to transfer the core state
orthogonality term in equation (2.237) from the OPW's to the one-electron
Hamiltonian H of the crystal. This transfer is done as follows. Consider
the core electrons. which we have hitherto taken jointly with the atomic
nuclei to form the cores, to be independent particles, just like the valence
electrons. This means that the potential created by the core electrons is no
longer included in the core potential Vc, but is added to the eEective o n e
+
electron potential 1 ' ~ 15-cof the electron-electron interaction. Because of
this reinterpretation of the oneelectron Hamiltonian H , its eigenstates now
also include core states, for which we have

H i c k ) = E,lck). (2.238)
The valence band eigenvalues E and eigenfunctions ~a
similarly satisfy the
eigenvalue equation

The expansion of ~$m


with respect to the OPW's reads

Removal of the core states from the expansion functions mandates changing
from the eigenfunctions $ato other functions v p i that have the same
expansion coefficients, but plane waves as expansion functions:

(2.241)

Surprisingly, the artificial wavefunctions ~$g~


are in fact eigenfunct ions of
a particular Hamiltonian H p s , having the same eigenvalue E to which the
eigenfunctions $aof H correspond. In fact, by applying of H to $I% one
immediately Ends that
2.5. Band structure 137

(2.242)

H P S= H + E ( E - E,)lck)(ckl . (2.243)
C

The Hamiltonian HPS is a well-defined linear operator, although it is non-


local and depends on the eigenvalue E itself. It is called a pseudo-Hamilto-
nian The $gk are termed pseudo-wavefunctions. Each eigenvalue of HP" is
simultaneously an eigenvalue of H , but the reverse statement does not hold.
While core levels E, are eigenvalues of H , they are not also eigenvalues of
HPS since the pseudo-wavefunctions of the core states vanish.
The pseudo-Hamiltonian HPS may be written in a form which clarifies its
meaning. To begin with one revokes the re-interpretation of the Hamiltonian
H , i.e. considers the core electrons no longer as independent particles which
+
enter the effective one-electron potential VH Vxc of the electron-electron
interaction, but includes them again into the atomic cores making them
contributors to the core potential V,. Then HPS takes the form

H P s = -p2+ v
2m
H + Vxc + Vc+ C ( E - E,)lck)(ckl . (2.244)
C

The last two terms in this expression jointly constitute the so-called pseu-
dopotential V,p"

(2.245)

The second term on the right is significant only in the core regions. There,
it is preferentially positive, indicating that it repels valence and conduction
band electrons away from the cores. One can show that it largely compen-
sates the variation of the core potential V, in these regions. Because of this,
the pseudopotential V,p" is relatively smooth throughout the cores. It can
he made even smoother if one exploits a property of V,p" which has not yet
heen discussed. We refer to the non-uniqueness of the repulsive part of V,p"
in equation (2.245). If there the bra-vectors (E - E,)(ckl in (2.245) are r e
placed by completely arbitrary functions (while keeping the ket-vectors Ick)
unchanged), the eigenvalues of the pseudo-Schrodinger equation (2.242) re-
main the same, only the pseudo-wavefunctions change. This freedom may be
used to make the pseudopotential still smoother, and also fulfill other require-
ments, such as, for example, norm-conservation of the pseudo-wavefunctions.
The smoothness of the pseudoptential makes it possible to restrict the plane-
wave expansion (2.241) of the pseudo-wavefunction to terms having small
reciprocal lattice vectors K. Consequently, the representation matrix of the
pseudo-Hamiltonian HPS is small and easy to diagonalize. This is the reason
138 Chapter 2. F k t m n i c structure of ideal crystals

that the pseudopotential method is very helpful in calculating valence and


conduction band structures of semiconductors.
In order to apply this method, the psendopotential must be known explic-
itly, of course, In principle it can be determined from the defining equation
(2.245),with core levels and wavefunctions taken from atomic calculations,
and with bra-vectors (ckl substituted by appropriate functions. In practical
applications, the pspudopotential V,p"i s replaced by approximate expressions
which range from empirical local pseudopotentials with adjustable param-
eters up to non-local pseudoyotentiah including core states of s-! p- and
d-symmetr y.
The pseudopotent,ial method is generally successful if the true valence
band eigenfunctions are also sufficiently smooth outside ol the core region.
This i s the c a ~ eas long as the valence band states are composed mainly
of atomic 3- and p-orbitals. If d-orbitals cont,ribute to these states in an
essential manner, i.e. if the d-electrons of the atoms are sigmficantly involved
in t,he chemical bonding of the cryshl, then the pseudopotential method
becomes problematic, because it,s main advantage in having the pseudo-
eigenfunctions built up from a relatively small number of plane waves, no
longer applies. Thus may occur in trhe rase of TII-V and 11-VI wmpound
semiconductors whose cations have flat d-levels as in the case of Zn, for
exmaple (see Table 2.2).
A completely different approach to solving the oneelectron Schrodinger
equation is taken in the so-called cell methods.

Cell methods

These methods are 3-dimensional generalizations of the method of match-


ing conditions usually employed in solving thr Schrodinger equation for the
square well potential and other 1-dimensional potentials. In cell methods?
one first determines linearly independent solutiona of the Schrodinger equa-
tion within a primitive unit cell for arbitrary energies. Forming Bloch sums
with them, one constructs the solution for the total crystal. The require
ments that these functions, and their first derivatives, be continuous at the
boundaries of the unit cells determines the energy eigenvalues and eigenfunc-
tiuns. This method suffers from the arbitrariness of t,lie choice of the unit
cells and the difficulty of satiseying the boundary conditions over the whole
surface, i.s. at an infinite number of points. The most natural choice of unit
cell is the Wigner-Seitz cell. A further development, of cell methods lies, in
a sense, in the mufin-tin methods.

Muf&n-tin methods

In the present method one delimits spheres around the atoms, and leaves
some empty space around them. The crystal then looks something like
2.5. Band structure 139

a muffin-tin, whence thc methods narnp. Within the spheres one uses the
spherimlly syuunelric potentials of the atoms, trnd in the surrounding regions
one takes the potential to be uniform.

In a special mufin-tin method, t,he augmented plane wave ( A P W)-method,


the solutions of the Schriidinger equation within the spheres are expanded
with respect to angular momentum eigsnfunctians. The radial parts of the
expansion are determined by the radial Schrijdinger equation. This is solved
for the various angular momentum quantum numbers numerically. Then, the
still unknown expansion coefficients are determined by the requirement that
t,he soliit,ions within the spheres mist join continuausly with the solutions
+
outside. The latter are plane waves of some wavevector k K. The func-
tions constructed in this way are called augmented plane wawes {APWs).
They are, of course not eigenfunctions of the oneelectron Schrdinger equa-
tion of the crystal, but m a y be taken as a basis for them. In contradistinction
to bhe basis functions used in matrix methods, the APWYstill depend on
the unknown energy eigenvalues. An rigenfunction expansion with respect
to APWs leads, as in matrix methods, l o B homogenwus set of equations
or the expansion coefficients. but the matrix elements are, unlike the Hamil-
tonian in matrix nielhoads, funcbiom of energy. However, the A PW-matrices
are in general smaller, as a consequence of the use of better adjusted ba-
sis functions. Often, a linearized energy dependence of the matrix element,s
yields useful results in the linearized APW o r LAPW method. If, in con-
structing the APWs, Gaussian functions are used instead of the angular
momentum eigenfunctions, one speaks of rnuff-tin orbitals (MTO s), upon
which the MTO and LMTO methods rest.

The different APWs are indexed by reciprocal lattice vectors K, in


addition to angular momentum quantum numbers. In another muffin-tin
method, called the Komnga-Kohn-Rostoker (KKR)-metho$, which takes ad-
vantage of the formal scattering theory of quantum mechanics in the Greens
function formulation, the expansion functions are also angular momentum
eigenfunctions between the spheres.

In band structure calculations for semiconductors, some of the methods


listed above are used more frequently than others. Among the ab initio
procedures, the pseudopotentid method combined with density functional
theory in its local approximation, and lately with the Greenss function
method, is particularly important. In addition, also the APW and LMTO
methods are used. mainly in their linearized forms. Of practical importance
among the empirical procedures are, above d l , the empirical versions of the
tight binding and of the pseudopotential methods.
140 Chapter 2. Electronic structure of ideal crystals

2.6 Tight binding approximation


In the nearly-free-electron approximation, the eigenstates of the one-electron
Hamiltonian of a crystal are represented by a superposition of plane waves.
The non-diagonal matrix elements of the periodic potential with respect to
these functions are treated as a small perturbation. Thus the eigenstates are
weakly disturbed plane waves in which lhe electrons are spread out almost
uniformly over the whole crystal. Such a distribution might be valid for
drctrons of tho conduction band, biit it does not correspond to the reality
one should expect for valence electrons if one considers the crystal to be
formed from previously isolated atoms. In free atoms, the valence electrons
are l o c a l i d at their respeclive atomic cores. Although this localization must
be partially breached in the crystal, in order for chemical bonding to take
place, an almost complete deloralization such as is assumed in the nearly-
free-electron approximation is not to be expected. This suggests a more
appropriate approximation which takcs atomic wavefunctions as the basis
set and treats the non-diagonal elements with respect to these functions
as small perturbations. This approximation is callcd a tzght b m d m g (TB)
appro.mmntzon The errors of this approximation are expected to be small
if the valence electrons of the crystal are well localized at the atoms. The
approximation of nearly free electrons will work very poorly in this case,
while it gives good results if the electrons are weakly localized, i.e. when
the tight binding approximation is not applicable. In this sense the two
approximations are complementary. Results which are obtained from both
these approximations may be considered to be independent of any particular
approximation, i.e. to be exact. The term exact here means within the
framework of the simplifications made earlier. One of these simplifications,
the approximation of frozen cores introduced in section 2.1, is particularly
important because it allows us to deal with only the valence electrons of the
free atoms, while the core electrons are incorporated in the atomic cores.
In this section we will develop the basic principles of the TB approxima-
tion. These principles will be applied to semiconductors of the diamond and
zincblende type, and it will be shown that the TB approximation not only
is capable of explaining the valence band structure of these crystals, but it
also provides insight into their chemical bonding and atomic structures.

2.6.1 Fundamentals
Atomic Orbitals

The basis functions of the TB approximation are the one-particle eigenfunc-


tions of the valence electrons of the free atoms, more strictly, of the atoms
composing the crystal under consideration. These eigenfunctions are called
2.6. Tight binding approximation 141

atomic oibatnls. The spinor character of the orbitals may be taken into ac-
count, but we will omit it here brevity. The mobt important property of
the orhit&, which turns out to be decisive for the TB approximation, is
their spatial symmetry. The latter i s determined by the symmetry of the
Hartree o i HartrecFock potentials of the atomic cores. These potentials are
isotropic if all core shells are fully populated by electrons. In the case of
atoms forming diamond and zincblende type semiconductors, this condition
is always satisfirul. Thus we may assume isotropic core potentials, and the
atomic orbitals of given energy eigcnvalues form basis sets of irrediicible rep-
resentations of the full orthogonal symmetry group. These representations
are characterizd by an angular momentum quantum number 1 which may
assume all non-negative integral values. The irreducible representation of a
+ +
given quantum number 1 is (21 1)-fold degenerate. The 21 1 basis func-
tions are distinguished by the magnetic quantum niimbcr 7n which takes all
integei values between - 1 and + l . The energy spectrum of the free atom
is degenerate with respect to m. Foi the hydrogen atom there is also a d c
generacy with respect to 1. This is not due to the spatial symmetry of the
potcntial but to its purr Coulomb form. Here this additional degeneracy
may not be assumed. For each value o l lhe quantum number 1 one has an
infinite set of different energy eigenvalues. These are distinguished by the
+
mdin quantum number n which may take all intcgcr values horn 1 1 to 00.
In this way the energy levels E,l of an atom depend on thc two yuanturn
numbers r~ and 1 , and the corresponding eigenfunctions dnlm(x)depend on
thrw quantum numbers r ~I ,, m. 'l'ht. eigcnfunctions &,jrn(x) can be written
as products of the radial wavefunctions &(I x I) and spherical harmonics
Km(xl I x I),

(2.246)
Here, it is assumed that the tttorriir core is located iit the coordinate origin
x = 0. A wavefunction with the quantum number I = 0 is called an s-orbital,
one with 1 = 1 a p-orbital, with 1 - 2 a d-orbital etc.
In order to represent the eigenstates of the valence electronb: of a crystal,
one needs, rigorously speaking, nll orbitals of the cores of its atoms since
only the totality of all orbitals forms a complete basis set in Hilbert space.
However, not all of these contribute in an essential manner. 'l'he largest
contributions are to be expccted from orbitals forming the valence shells of
the free atoms. Within the TB approximation one takes only these orbitals
into account, This corresponds to a perturbation-theoretic treatment of the
Hamiltonian matrix with respect to the atomic orbital basis; only matrix
clcmerits between valence orbitals are considered while those involving other
orbitals are neglected. For the elemental semiconductors of the fourth group
of the periodic table the valence shell orbitals are formed by the four ns-
142 Chapter 2. Electronic structure of ided crystals

and np-states with n = 2 for C , n = 3 for Si, n = 4 for Ge, and n = 5


for a-Sn. In the case of semiconductors composed of different elements, the
valrnce shell orbitals of the various atoms must be considered. For GaAs
that means the one 4s-state and the three 4p-states of Ga, and the one 4s-
and t h e Q-states of As. For GaP one bas, besides the abovementioned
4s- and lip-states of Ga, the one 3s-state and the three Spstates of P.
The valence shell orbitals used as basis functions need not, of course, be
populated by electrons in the case of the free atoms. For Si, for example, two
of the three porbitals are empty. Similarly, the eigenstates of the crystal,
which will be calculated later by means of the T B approximation, will also
not be completely populated. As this applies to quantum mechanics in
general, the eigenstates are candidates for posstble population. Whether
they are popdated or not. depends on the macroscopic state of the system,
e.g., on the temperature of the crystal
In the examples considered above the valence shells of the atoms are
formed by s- and p-states. This is the typical case for tetrahedral semi-
conductors composed of elements of the main groups 11, IV, and VI of the
periodic table, but it is by no means the only possibility, especially if one
also includes other material classes. For body-centered cubic metals such
as Cu and Ni, for example, the valence shells are formed by 3d- and 4s-
states. In section 2.1 it was already mention4 that d-states may contribute
to the valence shell of 11-VI semiconductors with heavy metal atoms such
as Zn. Here, we wiU exclusively consider semiconductor materials for which
only s- and p-states need to be taken as basis functions. The corresponding
spherical harmonics X,(X/ I x 1) are

Using these harmonics, eigenfunctions &oo, $n.cll,&lo, &i-i may be formed


according to equation (2.246). Instead of drill and d ~ ~ 1 - 1one
, may use the
linear combinations

The latter are also energy eigenfunctions because of the degeneracy of qbn1l
aiid 0 ~ 1 - 1 with respect to the magnetic quantum number m. For the sake
of uniformity, one sets
2.6. Tight binding approximation 143

Figure 2.16: Polar diagrams of the atomic s-and p- orbitals in Cartesian represen-
tation.

The eigenfunctions of equations (2.246) to (2.248) will be referred to as


spherical orbitals, and those of (2.249), (2.250) as Carte3ian orbitak. The
latter are visualized in Figure 2.16. The quantum numbers nlm of the spher-
ical orbitals, as well as n, 6 , z, y, z of the Cartesian, will be abbreviated by
a general index a. The orbitals considered above are orthonormalized, i.e.
one has
($a I $fZ) af,. (2.251)
If the atomic core is not located at the coordinate origin, as has been assumed
thus far, but at a particular lattice position R+G, the corresponding orbitals
will be denoted by &j~(x). They may be traced back to the orbitals &(x)
of atoms located at the origin by shifting their arguments in accordance with

Two different orbitals & f j f R ( X ) and & ~ R ( x ) with identical values of R and
R as well 8s of j and j , but different values of a and a, are also orthogonal
to each other. For R f R or j # j , i.e. for orbitals at different cen-
ters, no such orthogonality exists. Although the two orbitals are localized
in different spatial regions, and the integral over the product of the two, the
so-callcd ovcrlap integral, turns out to be relatively small, it may not be n e
glected because its influence on the energy eigenvalues is of the same order
of magnitude as the matrix elements of the Hamiltonian between orbitals
at different centers. The latter elements are essential because they are r e
sponsible for the bonding between atoms in a crystal and for the splitting of
the atomic energy levels into bands. The non-orthogonality overlap integrals
must therefore also be taken into account. This may be done directly, by
writing down and solving the eigenvalue problem for the crystal Hamilto-
nian in the non-orthogonal basis set of the atomic orbitals. This procedure
144 Chapter 2. Ektranic structure of ideal crystds

is, Iiowcvcr, quite inconvenient becaiise the matrix of overlap integrals has
to be calculated explicitly and diagonalized together with the Hamiltonian
matrix. It is more useful to employ a set of orthoganalizd orbitals by form-
ing suitable h e a r combinations of the q a j ~ ( x )Here,
. 'suitable' means that
the new orbitals should have t,he same spatial symmetries a s thc original
atomic nrhitrals Q U j ~ ( xbecause
), these symmetries are the essential proper-
ties that allow the matrix elements of the Ilamiltonian to be reduced to a few
const.ants. Consequently, the new orbitals must, likewise form basis sets of
irreducible representations of the full orthogonal group, each set heing char-
a c t w i d by a particular angular morrieriturri quantum miniber 1. That there
are indeed linear combinations of atomic orbitals possessing these properties,
forms the content of Liimdin,'<sthcowm (we, e.g., Slater, Koster, 1954). We
will iise this theorem below. The orbitals m U j ~ ( xwhich
)! thus far have been
taken as ordinary atomic orbitals, will henceforth be understood as atomic
orbitals in the sense of Ldwdin's theorem. The orbitals q a j ~ ( xare ) now,
of course, no longer given by the expressions [2.246) to (2.250), but their
indices n E nlm will retain the meaning they previously had for the ordi-
nary atomic orbitals. hlthough explicit expressions for the G w d i n orbitals
& j ~ ( x )can in principle be providd, they will not be given here since thcy
are nut, n m d d if one proceeds in thc manner to be discussed below. The
Lowdin orbitals arc, by definition, aIso orthogonal for different centers, such
that

(&!jrR( I O a j R ) = bo'abj'$R'R. (2.253)


With Ro as an arbitrary lattice vector, the d U 3 ~ ( obey
x ) the relation

(2.254)

Bloch sums. Schrodinger equation in matrix representation

Consider, again, the Schrodinger equation (2.75) of the crystal. As before,


the eigenfunctions of this equation are taken in the form of Bloch functions
pyk(x). To represent the Bloch type eigenfunctions by means of atomic
orbitals o a J ~ ( xit) . is convenient to transform the latter into Bloch type
orbitals puJk(x). This is done by means of the k-dependent unitary trans-
format ion
(2.255)

The sum over R extends over all lattice points of the periodicity region. By
means of (2.254) one may readily verify that

(2.2 56)
2.6. Tight binding approximation 145

which identifies the & j k ( x ) as Bloch functions of quasfwavevector k. The


4 , j k ( X ) are called Bloch sums of atomic orbitals. If one replaces k in &jk by a
wavevector which differs from k by a reciprocal lattice vector K then, because
of the relation exp[iK. R] = 1, the original wavefunction is reproduced
expect for an unimportant phase factor. Therefore, it suffices to take k in
the first B Z . Obviously, the use of the Bloch sums (2.255) automatically
puts us in the reduced zone scheme. The orthogonality of the Lowdin orbitals
& ~ R ( x ) results in the orthogonality of their Bloch sums 4 a j k ( X ) , such that

(#alkj I #ajk) 6aa6j1j6k1k. (2.257)


In the TB approximat,ion, the eigenfunctions C p v k ( x ) of the Schrodinger equa-
tion are written as linear combinations

cPvlc(x) = x ( J j k I c P v k ) c P a j k ( x ) (2.258)
jQ

of Bloch sums q 5 a j k ( x ) . Employing this representation in the Schrodinger


equation (2.75), we obtain

x(ajk I H I ajk)(ajk I V v k ) = Ev(k)(ajk I cPvk), (2.259)


ja

where the matrix elements of the Hamiltonian are given by the expressions

with

(aj0 I H I ajR) =
s ( xR - 5,).
d 3 x ~ ! J : (-x? ~ ) H ~ , I -

In deriving equation (2.260), the lattice translational symmetry of H has


(2.261)

been used. For a given wavevector k, the (ajk I H I ajk) form a square
matrix with a finite number of rows and columns, the same as the number
of different orbitals per primitive unit cell which were used for the represen-
tation of the eigenfunctions, i.e. number J of atoms per primitive unit cell
times number A of orbitals per atom. The integrals ( a j 0 I H I ajR) in
(2.261) describe the interaction of electrons in orbitals a and a, where one of
the orbitals belongs to the atom at ? and the other to the atom at R+ ?,I.
These integrals substantially decrease with increasing distance between the
atoms, so that it suffices in most cases to extend the sum on R in (2.260)
over the nearest, or if need be, the second nearest neighbor atoms only. Once
the integrah ( a j 0 I H I ajR)and, through them, also the matrix elements
(ajk 1 H 1 ajk) are known, the energy eigenvalues and eigenfunctions of
146 Chapter 2. Electronic structure of ideal crystals

the Schriidinger equation may be obtained by calculating the eigenvalues


and eigenvectors of the ( J x A ) x ( J x A)-dimensional IIamiltonian matrix
of equation (2.260). This is an easily solvable task which occasionally can
be treated analytically, but, can in any case, be done numerically. For each
k one has J x A eigenvalues and eigenvectors. They will be partially degen-
erate if symmetrical k-vectors are considered. If k varies over the first B Z ,
the J x A energy eigenvalues form J x A energy bands.

Simple example: cubic crystal composed of s-atoms

We will illustrate the above discussion with a simple model. Consider a


crystal having a primitive cubic lattice and 1 atom per primitive unit cell
( j - 1). The atoms are placed at the lattice points R, with - 0. The
set of atomic orbitals will be restricted to one ns-orbital only, such that
a = ns. With J = 1 and A - 1, a matrix of size 1 x 1 = 1 is obtained.
This the solution of thP eigenvalw problem is, in fact, trivial. In evalu-
ating the matrix element (nslk IH I nslk), we will terminate the R-sum
in expression (2.260) at the nearest neighbor atoms. The latter are located
at R1,z = (fa,O,O),R3,4 = (O,fa,O),Rs,s = (O,O,fa). For the matrix
elements occurring in equation (2.260), we use the notation

(nslO I H I nslO) = ens, (2.262)

(nslO I H I nslRt) = pns, t = 1 , 2 , .. . 6 . (2.263)


Here, we have employed the fact that, for reasons of symmetry, all six neigh-
bor atoms give rise to the same value of the integral (2.261). The value of ,BnS
depends on the overlap of the ns-orbitals localized at adjacent atoms. With
increasing distance between the atoms, &, approaches 0. The value of E,,,
in a crude approximation, may be identified with the energy of the ns-level of
the free atom. In terms of E,, and Pns, the matrix element (nslk IH I nslk)
of (2.260) is given by
6
(nslk I H I n s l k ) = ens + &sxeikRt.
(2.264)
t=l

It follows that the energy eigenvalues Ens(k) of the Schrodinger equation are

Ens(k) = ern + pns2 [cos(k,a) + cos(kya) + cos(k,a)] . (2.265)

It is of interest to further discuss the eigenvalues e,,(k), to gain insight


into the formation of energy bands. In this context, the main quantum
number n will be allowed to take not just one value, as previously assumed,
2.6. Tight binding approximation 147

Figure 2.17: Energy bands in


Tight-Binding ayproxirnat,ion for
the simple s-atom crystal described
ks P3s
in the text.

P 2s

PlS

-
-IT
D
0
Wavevector (O,O,k,) -
l-r
-
a

but several. This means that the valence electrons in the free atom do not
occupy only one 5 level, but several s-levels ens differing in the value of n. To
justify the application of the results obtained above in the present case, the
matrix elements of H between s-orbitals having different values of n must be
negligibly small. We assume this to be true. In addition, we suppose that ens
is negative for all n, and that the sign of pns alternates, such that for n = 1
it is taken to be negative, for n - 2 positive, for n - 3 negative etc. This
behavior reflects the differing numbers of nodes of thc atomic wavefiinctions
for different values of n. The absolute values of Pns are expected to increase
with growing n , corresponding to the larger values which the ns-orbitals
with larger n have at the nearest neighbor aloms. The separation between
adjacent energy levels cn8 should, however, always be larger than 4Pns.

If the above conditions are met, the energy band dispersion along the line
(0, 0, kz) of the first B Z of the simple cubic lattice under consideration, has
the form shown in Figurr 2.17. Each level ens of the free atom gives rise to
an energy band of the crystal. The width of these bands amounts to 4&.,
thus increasing with increasing n. The bands are separated by forbidden
energy regions. Their widths decrease with growing n. If the distance be-
tween nearest neighbor atoms increaees, then the parameters fins of equation
(2.263) decrease because of decreasing ovrrlap of the orbitals, as has been
148 Chapter 2. Electronic structure of ideal crystals

pointed out above. The same holds for the widths of the energy bands, they
become narrower if the distance between neighboring atoms grows. They
approach the discrete levels ern of free atoms as this distance becomes in-
finitely large, corresponding to ,&, = 0. In the latter case. electron states
with different values of the wavevector component k, have the same energy,
ens. These levels are, therefore, highly degenerate. If the infinitely remote
atoms again approach each other, then the ,&,-values become finite. because
of the onset of overlap of neighboring orbitals. Correspondingly, the degen-
eracy of electron states with different kz is removed. and the discrete levels
of the free atoms spread into bands. The TB approximation quite naturally
explains, in this way, how discrete energy levels of the free atoms transform
into energy bands of the crystal. The gaps between the bands occur nat-
urally in this approach, since the discrete atomic levels are separated by
energy gaps from the outset. This stands in contrast to the approximation
of nearly free electrons, where the occurrence of energy gaps calls for an
explanation. On the other hand, 'bandwidth', in the form of the infinitely
broad energy continuum of the free electron. is present at the outset in the
latter approach. The gaps induced into the energy continuum were seen to
arise because of the strong perturbation of plane wave states by the peri-
odic lattice array of atomic coTes fox wavevectors on the Bragg reflection
planes. Comparison of the two approximation procedures reveals the differ-
ence between the underlying concepts - the TB approximation emphasizes
the atoms and the short-range ordered complexes of the crystal. while the
approximation of nearly free electrons focuses on the crystal as a whole and
the long-range ordering of the atoms. Such comparison also shows that the
short-range and long-range order concepts are equivalent in the sense that
they result in the same characteristic features of the electronic structure of
crystals -both concepts predict the existence of energy bands separated by
gaps. Figure 2.18 depicts the manner in which the bands and gaps arise in
the two approximations.

2.6.2 TB theory of diamond and zincblende type semicon-


ductors
Semiconductors of the diamond and zincblende type are tetrahedrally coor-
dinate cubic crystals with two atoms per primitive unit cell and four valence
shell orbitals per atom, among them one s-orbital([ = 0) and three porbitals
( I = 1). The application of the TB method to this specid case is of particu-
lar importance. It will be developed in the present subsection. Only nearest
neighbor interactions will be taken into account because this introduces con-
siderable simplification and nevertheless gives results of reasonable accuracy.
For the p-orbitah. we choose the Cartesian form, so that the orbital index a
takes the values ns,nz,ny and n z . The Cartesian components I , y r z refer
2.6. Tight binding approximation 149

Figure 2.18: Illustrat,ion of the origin of energy bands and gaps in the enerw
spectrum of a crystal.

to the cubic crystal axes. The main quantum number n will be suppressed
below because it may assume only one value here for each of the t,wo atoms,
although not necessarily the same. As a first step we determine the 8 x 8-
Hamiltonian matrix for diamond type crystals. Later: it will be generalized
to zincblende type structures.

Hamiltonian matrix

We start with matrix elements bet,ween orbitals at equivalent atoms.

Matrix elements between orbitals at equivalent atoms

These are elements of the general form ( a j k 1 H 1 a'jk). In the case of


diamond type crystals. the two atoms J = 1 and J = 2 of the primitive
unit cell are of the same chemical nature, thus their matrix elements will be
identical. Since we are restricting ourselves to nearest neighbor interactions,
only the term with R'= 0 needs to be considered in the R'-sum of formula
(2.260). Thereby, ( a j k 1 H I a'jk) becomes independent of k. The eigen-
functions I s l O j and I zlO), 1 ylO), \ 210) transform, respectively, according
to the unit representation rl and the vector representation I'15 of the cubic
point group Oh. Since oh is the symmetry group of H , the matrix element
( s l k I H 1 s j k ) likewise belongs to the unity representation. According to
Appendix A this means that its value, in general, is non-zero, and we denote
150 Chapter 2. Electronic structure of ideal crystals

it by E ~ . In a crude approximation, es is the energy of the s-orbital of the


free atom.
The matrix elements between s- and p-orbitals at the same atom trans-
form according to the representation rlx I'l x r 1 5 = r15, which does not con-
tain the unit representation. Then, according to Appendix A, these elements
must vanish. The matrix elements between p- and p-orbitals belong to the
representation r 1 5 x I ' l x r 1 5 = X'1+r12+I'i5+I'b5 (see Table A.27). Here, the
unit representation occurs exactly once, which means that the p - p-matrix
contains exactly one independent constant. A more detailed analysis shows
that this constant corresponds to the three non-vanishing and mutually iden-
tical elements (210 I H I s10) = (y10 I H I y10) = (210 I H I 210) = E,.
Again, as in the case of c8, the value of cp is roughly the energy of the corre-
spondingp-orbital of the free atom. The non-diagonal p -p-matrix elements
must be zero according to the above symmetry analysis. Summarizing, we
have

(2.266)

Matrix elements between orbitals at non-equivalent atoms

In evaluating the matrix elements between orbitals at non-equivalent atoms,


i.e. elements of the general form (ujk 1 H I a'j'k) with j # j ' , the R'-
sum in (2.260) may be restricted to the 4 lattice points &,t = 1,2,3,4,
whose primitive unit cells host the 4 nearest neighbor atoms. With this
simplification the Hamiltonian matrix of (2.260) becomes

4
(ujk I H I a'j'k) = x e i k ' ( R t f 5 f - % ;(. )~ j IH
0 I U'j'Rt). (2.267)
t=1

The values of j and j' are complementary to each other because the nearest
neighbor atoms lie in the other respective sublattice. For j = 1 one has
j' = 2, and for j = 2 then j' = 1. We will restrict ourselves to the first case,
i.e. j = 1,j' = 2, because the second may determined from the first with
minor changes. The four nearest neighbors of a 1-atom are located in the
primitive unit cells at

R1 = 0, R2 = -al, R 3 = -a2, R 4 = -a3, (2.268)

(see Figure 2.19). The calculation of the matrix elements (a10 I H I a'2Rt)
between the different Cartesian orbitals and the four different values of Rt is
2.6. Tight binding appmxhation 15 1

24

Figure 2.19: Atom of sublattice 1 and its four nearat neighbors in sublattice 2.

somewhat laborious. It will be carried out in several (five) steps. In the first
step we determine the matrices (a10 I H I a2Rt) between spherical orbitals.

i) Spherical orbitals in the pair-related coordinate system

The spherical orbitals are related to a spherical coordinate system whose


z-axis is in the direction of the line connecting the central 1-atom and its
nearest neighbor atom 2t in the unit cell at & (see Figure 2.19). To dis-
tinguish between this z-axis and the crystal-related cubic z-axis, we denote
the former by 2. For each of the four next neighbor atoms, I has B different
direction. The matrix elements (a10 I H I a2Rt) between the so-defined or-
bitals are equal for all four nearest neighbors. The orbitals differ, however,
because they refer to different pair-related coordinate systems. Consider,
in particular, the neighbor atom 21 which lies in the same unit cell as the
central 1-atom. Here, R1 = 0. The corresponding Z-axis represents a %fold
symmetry axis of the crystal which contains three mirror symmetry planes.
In evaluating the matrix elements (Im10 I H I Irn20) one may therefore use
the fact that the Hamiltonian exhibits the symmetry of the point group C h
with respect to the Z-axis. This yields diagonality of these elements with
respect to the magnetic quantum numbers rn, m, so that

(Lm1O I H I Im20) = & , , , ( h l O 1 N 1 lm20) BmmrV~ppn


12 , (2.269)

The Cg,-symmetry of the Hamiltonian H holds with respect t o the nearest


152 Chapter 2. Eiectronic structure of ideal crystals

neighbor dirwtions. For the second-nearest neighbor directions, the symme-


try of H is smaller. Rwauw of that, the matrix elements (alk I H I a2k)
of equation (2.260) are, in general, non-diagonal with respect to m, m if the
R-sum is extended to the swond-warest or more remote atoms. The diago-
nality holds, however, in an approximate sense, which may be seen as foHows.
The periodic potential V(x) in H represents a sum &,,,, vji,(x - R - i y )
of potential contributions of all the individual atomic cores of the crystal.
The matrix elements of H between orbitals at different renterrs jR and JR
thereby decompose into bums over all centers JIR. The largest contribu-
tions will arise from t a m s where the center index jR coincides either with
1R or with JR. If one considers only such twecenter terms and neglects
all threecenter contributions, then the integrand of the matrix elements
(Im10 I H I lm20) has the full axial symmetry, so that these dements
become diagonal with respect to m, m.
The hermiticily of the Hamiltonian and the particular form of the wave-
functions in (2.252) lcsd t o the relation

(lrn10 I H I Im20) = (lm20 1 H I lm10). (2.270)


The three matrix elements

(0010 (2.271)

(0010

(1010
are independent of each other: and the two elements

(1110 I H I 1120) = ( i i i o I H I 1720) = v& (2.272)


are identicaL These elements are illustrated in Figure 2.20. This figure also
shows how the elements (In110I H 1 lm20) behave if the two atoms 1 and
2 are interchanged. One has

(2.273)

Since the atomic orbitals under consideration are those of bound states, and
since the eigenvalues of the Hamiltonian for bound orbitals are generally
negative, one may expect negative values for k<3n, Vm, and positive ones
for Ifaw, IT&,. Taking account of the strength of the overlap of the orbitals
2.6. Tight binding approximation 153

12
v s su
21
= vuG
-= vssu

21 21,
=V p p u = VPPT
-
=Vppcr

Figure 2.20: Illwtratiori of tight binding matrix elements.


154 Chapter 2. Electronic structure of ideal q y s t d s

in Figure 2.20, one can conclude that the absolute values of these elements
should obey the relations

I VPPO I I v, I I VSSU I >I V, I. (2.274)

These expectations are in fact valid in most cases.

ii) Cartesian orbitals with respect to the pair-related coordinate system

The matrix elements (lm10 I H I ImflO)of H with respect to spherical or-


bitals calculated above will be used to derive, in the second step, the matrix
elements of this operator with respect to the Cartesian orbitals. The z-axis
of the Cartesian coordinate system is taken t o be the same as that of the
spherical coordinate system used above. This means that the z - ax is points
in the direction of the connecting line between atom 10 and atom 21. The
corresponding 2- and y-axes lie in the plane normal to the z-axis, apart from
an irrelevant rotation about this axis. The pair-related Cartesian coordinate
system thus defined differs from the formerly introduced crystal-related sys-
tem which is given by the three cubic crystal axes. The coordinates in the
pair-related Bystem will be denoted by 2 , 5,2. The pair-related ;-orbital co-
incides with the 8- orbital with respect to the cubic-axes system. In terms
of this notation, the matrix elements of H between Cartesian orbitals in
the pair-related system read (310 I H I 3.20),(3.10 I H I 120),(3.10I H I
520), (3.10 I H I .520),(110 I H I 120), (110 I H I 520) etc. Since the C a r t e
sian orbitals are defined in terms of spherical orbitals by equations (2.249),
(2.250), the matrix elements between Cartesian orbitals can also be related
to those between spherical orbitals. The corresponding relations are given
below. Elements which are complex conjugates due to hermiticity of the
Hamiltonian, such as (310 I H I 220) and (120 1 H I ;lo), are listed only
once. The relations read

(310 I H I 3.20) = VSScr,


(3.10I H I 520) = (3.10 H I 520) = 0,
(3.10 I H I 220) = VSF,
(110 I H 1220) = (510 I H 1520) = Vw,
(510 I H 17/20) = (g10 I H 1.520) = (210 I H 1120) = 0,
(El0 I H 1.220) = Vppu. (2.275)

To develop the representation of the Hamiltonian matrix (2.267), we need


the elements (a10 I H 1 a2Rt) of H between Cartesian orbitals which refer
to the three cubic crystal axes rather than to the pair-related ones. We
determine these elements in the third step.
2.6. Tight binding approximation 155

iii) Cartesian orbitals in the crystal-related system

To this end, the Cartesian orbitals 2,y, z related to the cubic-axes system
must be expressed in terms of the pair-related Cartesian orbitals ?,c,Z.
To determine this relation, we consider a rotation which transforms the
crystal-related axes system into the pair-related one. The transformation is
characterized by Euler angles $, 6 and 'p. Since the orientation of the pair-
related system is defined only up to an arbitrary rotation about the Z-axis,
the Euler angle 'p may, without any loss of generality, be set equal to zero. As
noted in Appendix A, the rotation matrix A of equation (A.31) transforms
the coordinates before rotation into the rotated one. The basis vectors are
transformed by the transposed matrix, which, in the present case of rotation,
is also the inverse matrix. Since the 2 ,y, z and ?,y, Z are understood here
in the sense of basis vectors, we have

(:) cos @
= (sin@
0
- cos 6 sin $
c o ~ e c o ~
sin 0
sin 0 sin $
-sinecos$)
~
cos e
(i). (2.276)

Below, we will see that the direction cosines (PI,q1, 71) of the pair-related
2 -axis with respect to the crystal-related cubic z-axis play an important
role. These are the elements of the third column of the rotation matrix in
equation (2.276), i.e.

p l = sin 0 sin $, q1 = - sin 6 cos $, r1 = cos 0. (2.277)


Using equations (2.275) to (2.277), the matrix elements in the crystal-related
system are evaluated as One obtains
156 Chapter 2. Electronic structure of ideal crystals

These relations are valid for matrix elements involving the two nearest neigh-
bor atoms belonging to the same unit cell at R = 0. From these elements
we may determine the elements with the nearest neighbor atoms of different
unit cells. This will be done in the next step.

iv) fourth step.

The vectors pointing from the central 1-atom to the four nearest neighbor
atoms will be denoted by dt, t = 1,2,3,4, such that dt = Rt +
- TI.
For the diamond structure, the two sublattices are displaced with respect to
each other by the vector (a/4)(1,1,1). Using equation (2.268) for Rt and
the explicit form of the primitive lattice vectors of the face-centered cubic
lattice, we obtain

a - a - -
dl = a(l,
4 4
U
1, l ) , d2 = -(1,1,1),d3 = -(l, 1,I), d4 = -(l, 1,l).
4 4
(2.279)

These relations determine the direction cosines ( p t , qt, r t ) of the connecting


lines between the central 1-atom and the nearest neighbor atoms 2t. In par-
ticular, (PI,a ,7.1) = ( l / f i ) ( L 1,1),( P Z , 4 2 , 7 2 ) = (l/fi)(LI,l)etc. The
unknown matrix elements (a10 I H I a2Rt) for t = 2,3,4 follow from the
elements (a10 I H 1 a2R1) in equation (2.278) by replacing the direction
cosines (PI, m, ri) by ( p t , qt, r t ) in them.

v) fifth step

The elements (a10 I H I a2Rt) determined above are used to calculate the
k-dependent matrix elements (alk I H I a2k) between Bloch sums. For
complex conjugate elements, again, only one relation will be given. With
the help of equation (2.267), we obtain

4 4
(slk I H 1 s2k) = ~ e i k - d f ( s 1I0H I s2Rt) = V,,,xeik.dt, (2.280)
t=l t=l

4 4
(slk 1 H I 22k) = x e i k . d t ( s 1 0 I H 1 22Rt) = Vswxeik-dtpt.
t=l t=1
2.6. Tight binding approximation 157

4 4

The combinations of matrix e1ement.s in these expressions are commonly


abbreviated by

The seven different t-sums which enter t,he matrix elements (2.280) may be
reduced to just four because of the obvious relations ptpt = r t , p t r t = qt. and
qtrt = p t . The four independent sums are
158 Chapter 2. Electronic structure of ideal crystals

9l(k) = eikdi + ,ik.dz + ,ik.d3 + ,ik.d4, (2.282)


g2(k) = eik.dl + .ik+d2 - .ik.d3 - .ik.d4
9 3 0 4 = .ik.d~ I .&.dz + .ik.ds - ,ik.dn:
9 4 0 4 = .ik.dl - - e ik-dz - ,ik.ds + ,ik.d4

Finally, we c an write down the Hamiltonian matrix (ajk I H I ajk) between


Bloch sums in explicit form. Arranging the eight basis functions 1 ajk) in
the sequence I s l k ) , I z l k ) , I y l k ) , I z l k ) , I s2k), I 22k), I y2k), I z2k), this
matrix is given by

Band structure. Empirical TB method

To obtain the band structure of diamond type crystals, one has to calcu-
late the eight eigenvalues El(k),Ez(k), . . . , Es(k) of the Hamiltonian matrix
(2.283) at the various points k of the first BZ. The k-dependence of the
eigenvalues stems from the factors 91, 92, g3,g4 in (2.282), which are univer-
sal functions of k. The constants r,, e p , E,,, E,, E,, and E,, are related to
material properties. In principle, these constants can be calculated from the
defining equations (2.280) and (2.281) and the Hamiltonian matrix elements
(2.261) between atomic orbitals. In quantum chemistry one often proceeds
in this way. However, one also may forego the calculation of these Hamil-
tonian matrix elements and considers them as empirical parameters. Their
values may obtained as follows. First of all, one calculates the eigenvalues
of the Hamiltonian matrix (ajk I H I ajk),at at least 6 special k-points as
functions of the unknown parameters cg, epr E,,,E,, E,, and E,. Then one
measures the band energies at these points 01calculates them by some other
2.6. Tight binding approximation 159

1 1 1 1 1 1 1;1
Table 2.7: Tight binding matrix elements for diamond type semiconductors (in
eV). (After Harrason, f9RO.)

~i 6s EP VsSa V v vma
-17.52 -8.97 -4.50 5.91 10.41 -2.60

-13.55 -6.52 -1.93 2.54 4.47


Ge -14.38 -6.36 -1.79 2.36 4.15 -1.04

Table 2.8: Universal inter-atomic TB parameters. (After Majewski and Vogl,


1987.)

mi -1.38 1.92 1.68 2.20 -0.55

method. Equating the TB eigenvalues with those measured or known from


other calculations, one obtains a system of equations which determines the
unknown parameters uniquely. There are other variants of this procedure.
For instance, the intra-atomic matrix elements t 3 and tp may be identified
with the atomic s- and p-energies, leaving only the inter-atomic constants
E,,, ESP,Eppand Eppfor the fitting procedure. Instead of EsS,E,, E , and
E,, one often takes the inter-atomic matrix elements V,,,, Vspo,V,, and
V,, as independent parameters. Together with c3 and e p , they are referred
to as tight binding parameters. (see Table 2.7). Once these are determined,
one is able to calculate the band structure at all k-points. The whole pro-
cedure may therefore be considered to be an interpolation of band energies
between special points of the first B Z . It is called the empirical tight bind-
ing (ETB)-method. This method can also be applied to the calculation of
the electronic structure of perturbed semiconductors, which will be treated
in Chapter 3. In this case it serves as an extrapolation method, because it
helps to extrapolate from the electronic structure of the ideal crystal to that
of the perturbed one.

Inspecting the inter-atomic TB parameters listed for the various dia-


mond type semiconductors in Table 2.7, it may be seen that, within an
appropriate error limit, each of these parameters may be expressed in terms
of the material-dependent nearest neighbor distance do of the crystal, and a
160 Chapter 2. Electronic structure of ideal crystals

Table 2.9: Intra-atomic TB parameters (in eV) to be used in conjunction with the
universal inter-atomic TB parameters of Table 2.8. (Source same as in Table 2.8.)

universal material-independent factor qltm. This approximate relation is

(2.284)

signifying that the inter-atomic matrix elements of the Hamiltonian scale


with the inversesquare of the nearest neighbor distance. The factor ( f i 2 / m =
)
7.62 eVA2 in equation (2.284) has been jntroducd in order to make the
universal factors qzpm dimensionless. The nrarest neighbor distances dg fol-
low from the cubic lattice constants a in Table 1.2 by ineans of the rela-
tion do --- (fi/4a). Ultimately, the empirical do'-dependence in equation
(2.284) originates from the kinetic energy operator of the crystal Harnilto-
nian (li'royen, Harrison. 1979). The eigenstates of this operator are plane
waves and its eigenvalues are proportional to the inversesquare of the wave
length of i t s eigenstates. As a unnsequmce of this, the empty lattice band
structure scales with the inverse square of the latlice constmt. The TB band
structure follows by diagonalizing the T H Hamiltonian (2.283), resulting in
c l o d analytical expressionh for the energy hand lev& at special k-points
of the first BZ. If these expressions are identified with the empty lattice
band levels at, the 5ame k-points, respectively, linear eqiiations for the TB
parameters ctre obtained. Their solutions stale with the inverse-square of the
lattice constant a because the empty lattice band levels do so. hlormver,
numerical valiies for the universal TB parameters q ~ follow ' ~from these
equations. They are close t o the values listed in Table 2.8 which have been
derived from more precise band structure data (the meaning of figpa wiU be
explained below in connection with zincblende type semiconductors).
The intra-atomic matrix elements e g and c p corresponding to the set of
universal inter-atomic TB parameters ql+, in 'lable 2.8 are shown in Table
2.9 for a series of atoms forming diamond and zincblende type semiconduc-
2.6. Tight binding approximation 161

tors. They represent atomic s- and p-level energies which deviate somewhat
horn the energy levels given in Table 2.1.
A k-point of prtictilar interest is the 32 center k = 0. Here, the
eigenvalues and eigenfunctions of the Hamilton matrix (2.283) may br o b
tained ~ I closed
I analytic form. The results me the following 8 eigenvalues
&(a). i = 1,2,. . . , 8 .

The components of the corresponding eigenfunctions I EiO) read as follows:

1 E10) = (l/&)(l,o,o, 0,l.O. 0, O), 1 EzOj = ( I / f i ) ( O , 1,0, o.o,T, o,oj,

1 . ~ ~-0 ()i / ~ ) ( 0 , o , i , 0 , o , O , i , o ) , 1 E40) - (i/&j[o,o,o, i,o.o,o,ij,


I R 5 O ) = (1/&)(1,0, o,o, 1,0, 0, 01, I EGO) (1/&)(0, 1,0,0,0.1, o,o),
I E ~ O-) (1/&)(0,0, LO, o,o, i , o j , I E ~ O-) (i/fi){o,o,0, I, 0.0, o,i).
(2.286)
Since cP is negative and EZzpositive, the triply degenerate level 6 2 ( 0 ) =
E3(0)= E 4 ( 0 ) lies below the triply degcneratc level Rs[0)= Fe(O) = E 7 ( 0 ) .
The constants E~ and E,, both have negative values, therefore the El(O)-hvel
is lower than the Es(O)-level. Owing to the fact that the atomic s-energy
eS lies below thp pencrgy cp7 the pigPnvahx9 E l ( 0 ) is also smaller than the
deeper of the two Zriyly degenerate levels. This means that E l ( 0 ) is the
deepest of the foul levels. Thm+nrP, the energetic ordering of the lwels is
determined by the relatiom

El@) < ER(O), (2.2~17)


E l ( 0 ) < E z ( 0 ) - E3(0) = E4(0) < E5(0) E s ( 0 ) - &;7(0).
For the pmitinn of the &(O) lcvcl, there are still three pussibilitks (impor-
tant malerials to whirl1 the three possible cases apply are listed alongside
the cases), namely :

i) &(0] > &(O) > E s ( 0 ) - for C and Si,


ii) E3(0) < & ( 0 ) < & ( 0 ) - Ge,
iii) &(O) < &(0) < &(o) ru-sn.
~

These relations Inem that the E'c;(O)-levelmoves down with respect to the
other two levels as the size of the atoms increases.
162 Chapter 2. Electronic structure of ideal crystals

It turns out that the ordering of the eight energy bands at k = 0 remains
the same over the entire first BZ. This is important because the positions
of the energy bands relative to each other determine the likelihood of their
population by electrons. As already mentioned at the beginning of this
section, not all of the bands are expected to be populated, just as the s- and
p-levels of the free atom whose orbitals were used as basis functions were not
completely filled. Electron population of the ground state of the crystal, i.e.
at temperature T = 0 , may be obtained as follows:
For a simple band, each k-value corresponds to 2 eigenstates of opposite
spin. For a periodicity region of volume Q = G3Ro, the first B Z contains,
as does every primitive unit cell of reciprocal space, G3 allowed k -values
(see section 2.3). A simple energy band therefore has 2 x G3 states. Four
such bands are necessary to host the (2 x 4) G3 = 8G3 valence electrons
of a periodicity region. In the ground state of the crystal, therefore, the
four lowest bands are populated, and the four highest bands are empty.
This means that in the case of C , Si and Ge, E l ( k ) , E2(k), E3(k), E4(k) are
the populated valence bands, and Es(k),&(k), E7(k), Es(k) are the empty
conduction bands. For a - Sn, El(k) and &(k) form valence bands, together
with two of the three bands E2(k), E3(k), E4(k). The remaining band of the
three is a conduction band. Since it is degenerate at k = 0 with the highest
valence band, the energy gap of a-Sn vanishes.
The above band assignment allows us to determine the symmetry of the
valence and conduction band states at r.

Symmetry of valence a n d conduction band states at r


We know that degenerate eigenst ate8 of the crystal Hamiltonian having the
same energy value form a set of basis functions for an irreducible repre-
sentation of the small point group for the wavevector k. For k = 0 this
group coincide8 with the full point group of cquivalent crystal directions,
whch, here, is oh, The dimensions of the irreducible representations are
the same as the degrres of degeneracy of the corresponding energy lev-
els. Therefore, the eigenfunctions for & ( 0 ) and &(O) each belong to 1-
dimensional representations, and those of E2(0) = &(o) = E4(0) and
E s ( 0 ) = E s ( 0 ) = E r ( 0 ) each belong to 3-dimensional representations. Ac-
cording to equation (2.286), the deepest valence bend level Ei(0) ha5 the
eigenfunction (l/fi)[Is10)+ I s20)]. In order to determine its transfor-
mation properties under the operations of the point group Oh, it is useful
to decompose oh into two parts, firstly, the tetrahedron subgroup I>con-
taining only elements which are not involved with an exchange of the two
sublattices 1 and 2, and secondly, the remainder of oh which is composed of
all elemfxttb of T d multiplied by the inversion. Each of the elements of the
second part of 01, exchanges the two sublattices. For brevity, we will term
2.6. Tight binding approximation 163

the latter elements 'exchanging', and the former ones 'non-exchanging'. The
eigenfunction (l/&)[ls 1 0 ) + 1 sZO)] for E l ( 0 ) transforms into itself under
the action of both types of elements, thus it belongs to the unit representa-
tion rl. For the eigenfunction (l/fi){l sl0)- I s 2 0 ) ] of the Es(O)-level, the
transformation into itself occurs only under the action of non-exchanging el-
ements, while a factor -1 is generated in the case of exchanging elements. It
follows from the character table of the irreducible representations of Oh given
in Appendix A, that this transformation corresponds to the representation
r;.
The upper valence band level E z ( 0 ) = E3(O) = E4(0) possesses the
three eigenstates ( ~ / f i210)-) [ l 1 220)],( l / f i ) { Yloj-
l I Y ~ o )(1/]&)[1.
~ 1 0 ) - I d o ) ] . Under the action of the non-exchanging elements of oh.
these functions transform like vector components. Inversion, which is part
of the exchanging elements. reverses the sign of the vector components, and
the exchange of the two sublattices also reverses the sign of the whole eigen-
functions. The three eigenfunctions therefore transform as they w o d d under
the action of the corresponding non-exchanging elements. The character ta-
ble of the irreducible representations of Oh in Appendix A shows that this
transformation is characteristic of the representation J&. Similarly, one finds
that the eigenstates (1/&)[1x10)+ I d o ) ] ,(1/&)[1 y 1 0 ) + I Y ~ O ) ](, 1 / f i ) [ I
zl0j-t I ZZO)] belonging to the eigenvalues E s ( 0 ) = E e ( 0 ) - E7(O), trans-
form according to the irrcduciblc representation rl;.
We summarize the results of our TI3 band structure calculations as fol-
lows: For crystals which have the diamond structure, i.e. for C, Si, Ge,
as a rule, the highest valence band is 3-fold degenerate and belangs to the
irreducible representation r&, of the point group oh. The lowest conduction
band at r exhibits either a similar %fold degeneracy, in which case it be-
longs to the repraentation f 1 5 (C, Si), or i t is non-degenerate and belongs
t o the representation I 'i (Ge). For LY - Sn, the F2-band lies below the I'g
-band, which in this way, is partially a conduction band. These results are
illustrated graphically iu Figure 2.21.

Extension t o semiconductors of zincblcnde t y p e

Band structures of diianioand type semiconductors calculated by means of


the empirical TB method reflect the essential features of the real valence
bands of these materials quite well. In order to apply the TB approximation
to semiconductors having the zincblende structure, the Hamiltonian matrix
[Z.ZS3) must be modified as follows. Firstly, it has to be recognized that the
matrix elements E~ and between orbitals at the mme center depend on
whether the center is an atom of chemical species 1 or 2. This means that
two different s- and p-energies have to be inserted into the two 4 x 4 diagonal
blocks of the matrix (2.283);ti,t i in the block at the upper left, and e ; , :t
164 Chapter 2. Elect,ronic structure of ideal crystals

r, - 6, - r;, - conduction

6, - r, - r, - bands

r;, - valence

6 - bands

diamond structure zinblende structure

Figure 2.21: Ordering of energy bands at the center 1 of the fint B Z for semicon-
ductors of diamond and zincblende type.

in that at the lower right. Secondly, the relation V$$ -V:$ G -V3w of
equation (2.273) cannot be used in the zincblende case because it rests on
the chemical identity of lhe two atoms of A unit cell. For the matrix element

z;l 2 , the s-state belongs to a 1 atom, and the p-state to a 2-atom, while for
the s-state belongs to a 2-atom, and the p-state to a 1-atom. Therefore,
thr matrix element, V& is given by an independent constant -espc
rather
than by -V:$, (the parameters iSwin Table 2.8 are associated with csF by
meany of equation (2.284). The other excliangr relations in (2.284) remain
valid because they refer to matrix elements whose orbitals at the two different
atoms belong to the same state. With these two changes, in equation (2.283)
the TB Hamiltonian matrix for zincblende type semiconductors becomes
2.6. Tight binding approximation 165

with ESP= (l/fi)c&,. TTsing this matrix, the band structure and the
eigenstates of zincblende type semiconductors may be calculated. YIPspa-
tial symmetry of the eigcnstatcs a t thr R Z rentpr is similar to thtlf, which
was found for diamond type crystals above, except that the point group Ott
has to he r ~ p l a r dby the tetrahedron group T d . The wprPsPntaliun ri, of
Oh thereby becomes the representation r15 of T d , I 'i is replaced by rl, and
r1.5 remains ri5. In contrast to diamond type crystals, one has practically
only one energetic ordering of the conduction bands here - the rl-band being
the deepest.

2.6.3 sp3-hybrids, total energy and chemical bonding


Once the bald stnicture is known, the total energy of the valence electrons
ran be calculated. Formula (2.54) of section 2 . 1 indicates how this may be
done for the ground state of the crystal. One hw to sum the energy levels of
all valence electrons! and subsrquently remove the doubly counted electron-
electron Coulomb interaction energy from this sum. The total energy of the
vakncr elwtrons of the crystal or, strictly speaking, its deviation from the
total energy of the valence electrons of the free atoms which were brought
together to form the crystal, defines the energy gain due t o chemical bonding.
known as coh.esion energy of the crystal. This definition is reasonable because
the valence elect,rons are the only parts of the atoms whose stat= change
when the crystal is formed.
In order to obtain the total e n e r a in closed analytic form, one needs
explicit mathematical expressions for the valence band energies st. all paints
k of the f i d BZ. However, the TB approximation in the form developed
above produces such expressions only at particular symmetry points. To
find them everywhere, one must introduce further simplifications. A starting
point for this is a formulation of the 'I'B approximation which employs certain
linear combinations of 8- and p orbitals as basis functions, rather than the
atomic orhihals of definite angular momentum quantum numbers I ! which
were used above. These linear combinations are callcd sp'--hgrhr%d orbitals.
In contrast to the atomic orbitals, they are not energy eigenfunctions of the
atoms.
sp3-hybrid orbitals
The four hybrid orbitals I h~lR), I h z l R ) , I hslR),I hslR) of a 1-atom in
the unit cell at R are defind by the equations
166 Chapter 2. Electronic structure of ideal crystals

23 23

23 23

Figure 2.22: Illustration of the wavefunctions involved in chemical bonding in


tetrahedrally coordinated semiconductors: sp3-hybridorbitals (a, b), bonding or-
bitals (c) and anti-bonding orbitals (d).

1
I h31R) = 5 [dslR(X) - # z l r t ( X ) -tb y d x ) - $zlR(X) 1 1

1
I h d R ) = 5 [ $ s ~ R ( x-) 4 z d X ) - &lR(X) + d'.zlR(X) 1
In Figure 2.22 the probability distributions of the four sp3-hybrid orbitals are
shown in the form of polar diagrams. The orbitals resemble clubs pointing
to one of the nearest neighbor atoms in the sublattice 2, e.g., I hllR) to
atom 21, 1 h z l R ) to atom 22 etc.
2.6. Tight binding approximation 167

Similarly, the four sp3-hybrid orbitals I h12R),I h22R), I h32R), I h42R) at


a 2-atom in the unit cell at R, are defined by the relations

These orbitals point to an atom of type 1. The following four orbitals are
directed to the 1-atom in the unit cell at R: orbital 1 h12Rl+ R) at atom 21,
I h 2 2 R 2 f R) at atom 22, I h32R3+ R)at atom 23,and I h42&+ R)at atom
24. Hybrid orbitals at the same center are orthonormalized with respect to
each other, as the s- and porbitals from which they are constructed. If the
latter are understood in the sense of Liiwdin, orthogonality also holds for
hybrid orbitals at different centers. For each unit cell there are 8 associated
hybrid orbitals 1 &jR)with t = 1,2,3,4and j = 1,2, as there are 8 atomic
orbitals for the two free atoms of a unit cell. From the hybrid orbitals of a
given sublattice one may form Bloch sums I htjk) by means of the relation

(2.291)

in complete analogy to the Bloch sums q b ~ involving


k atomic orbitals in equa-
tion (2.255). Due t o the definitions (2.289) and (2.290) of the hybrid orbitals
in terms of atomic orbitals, the Bloch sums I h t j k ) may be thought to arise
from the Bloch sums # a 3 of ~ atomic orbitals by means of a unitary transfor-
mation. The same holds for the Hamiltonian matrix ( h t j k I H I htrg'k) with
respect to the basis I htj k); this matrix may be understood as arising horn
the above mentioned unitary transformation of the known Hamiltonian ma-
trix ( a j k I H I a'j'k) with respect to the basis daJk. As such, it has the same
eigenvalues and eigenfunctions as the original matrix. The implementation
of the TB method by means of hybrid orbitals instead of atomic orbitals
having a definite angular momentum quantum number is therefore nothing
but the solution of the same eigenvalue problem in another representation.
This statement holds, of course, only as long as equivalent approximations
are made in the two representations. However, the hybrid orbital represen-
tation is well s u i t e d for further approximations, beyond those already made
earlier. These approximations facilitate the derivation of simple analytical
168 Chapter 2. Electronic structure of ideal crystals

expressions for the eigenvalues and eigenvectors which may be used to ex-
plicitly determine the total energy of the valence electrons of a crystal in
closed analytical form.

Hamiltonian in hybrid-orbital representation

Here, we describe the most important additional approximation available in


hybrid orbital representation, which utilizes the fact that those particular
hybrid orbitals at nearest neighbor atoms which point toward one another
will also overlap one another more strongly than all others. In consequence
of this, the Hamiltonian matrix elements ( h t l R I H I ht2 R
t R) between +
these orbitals will be the largest. Approximately, one need only consider
these elements, while all others may be neglected. The former elements do
not depend on t and R, as can be seen from the explicit expressions for the
hybrid orbitals (2.289), (2.290). Their common value, denoted by V2, may
be determined from (2.289), (2.290) and equations (2.271) to (2.273) as

The corresponding matrix elements (h+lk I H I ht2k) between the Bloch


sums of hybrid-orbitals follow from VJ by multiplying this quantity with the
factor
et eak.dt
~ (2 293)
In the casc of Hemiltonian matrix elements between hybrid orbitals at the
same center, one has to distinguish between diagonal and non-diagonal ele-
ments. Thc non-diagonal elements ( ~ Q 1RIf 1 h t , ~ R have
) thc samc value
for all R and orbital quantum numbers t , t'. For the common value V1 one
finds by means of (2.289), (2.290) the expression
1
Vl 5 ( k t j R 1 H 1 ht,jR) - - ( c S - E ~ ,)
4
t # t' . (2.294)

with e3 and ep defined in equation (2.266).


Since only nearest neighbor atoms are considered, the rrvult is the same
as for the matrix element ( h y k I ZI I h t f jk) between the corresponding Rlocli
sums. An analogous result holds for the diagonal elements (&jR 1 H 1 ht3R)
at the same center. Their common value ctt is
1
~h (htjRI H I h y R ) = -(~g -t- kp). (2.295)
4
Again, this result is the same as for the matrix elements ( h u k 1 H 1 h u k )
between the corresponding Uloch sums. Numerical values for th, along with
values of Vi and VL, are listed in Table 2.10.
2.6. Tight binding approximation 169

In writing down the Bamiltonian matrix (htjk I H I hrjk) in the hybrid-


orbital representation, we arrange the eight basis functions in the sequence
I h l l k ) , I hzlk), I h s W , I h 4 W , I h l W , I h Z W , I h32k), I h 4 W . Then
the matrix ( h t j k I H I httjk) takes the form

. (2.296)

This matrix is also known as the Weare-lhorpe Uarniltonian. Remarkably,


its 8 eigenvalues can be obtained in closed analytical form. Denoting them
by E,b, E:, i = 1 , 2 , 3 , 4 ,we havve

Table 2.10: Hybrid matrix elements calculated from the TB parameters in Table
2.7 (in eV).

-8.27 -1.76 -2.98


Ge -8.37 -2.01 -2.76
170 Chapter 2. Electronic structure of ideal crystals

r X
wovevector

Figure 2.23: Evolution OF the energy bards of semiconductors with the diamond
slructura within the TB approximation The right-hand part shows the band struc-
t,ure of Si calculated by meam of equations (2.297).

Here gl(k) is the structuredependent factor d e b 4 in equation (2.282).


The actual positions and k-dispersions of the bands (2.297) are deter-
mined by the parameters th, 11 and Vz. Using the values for S i given in Table
2-10, one obtains t h bnrrd
~ structure shown in trhf right hmd part of Figure
2.23. In this regard. the 4 bands indicakl by b lie below the energy gap,
arid thP 4 bands indicated by u lie above it. This means that in the ground
state of t h e crystal. the &lands are fully populated by electrons, i.c. they
form the valence bands, and the a-bands are complekly empty, thus they
are the conduction bands. Essential features of these relationships are the
negative sign of V2, and the validity of the magnitude relation I V1 1>1 2 I
between the absolute values of and Vz. Figure 2.23 also illustrates how
the different bands emerge from the atomic s-and p-1evrlb due to the two
interactions V1 and V z . Roughly speaking, Vz determines the distance be-
tween the cpnters of gravity of the valence and conduction band complexes,
and V1 the width of thpse bands.
Now wc return to the main goal of this subsection. the ralcuIatjon of the
total energy of the crystal. To accomplish this, the energy values of the four
valence bands E,b(k)determined above must be summed over all i and k. It
2.0'. Yi'ghht binding approximation 171

turns out that this task may even be carried out analytically if a suitable
additional approximation, the so-called bond orbital approsirnation. is made.

Bond orbital approximation

To introduce this approximation, we review the already solved problem of


diagonalizing the Hamiltonian (2.296). but procwd in a sompwhat different
way, The diagonalization will now bt. carried out in two steps. In the first
step, the eigenvalues and rigenvectors of the matrix (2.296) are calculated
without taking account of the VI-terms, i.e. temporarily setting Vl to zero.
As may he scen from formula (2.297). this leads to the reduction of the
enerffi bands to the two dispersionless levels givw by

5k) belonging t o q, have the components

bik) ( l / f i ) [ l ,D,O,O,e;,O, O,O), (2.299)

(htlk 1 b4k) = ( l / J z ) ( O , 0: O , l , O , O , O , e : ) ,

and the 4 eigenfunctions I atk) belonging to E, have the components

(ht2k I agk) = (l/&)(O,O; 1,0,0,0: -ez, 0),

(ht2k I aqk) = ( l / & ) ( O , O , O , 1.0,0,0, -.;I.


Each of the eigenfunctions I btk) and 1 atk) is a linear combination of Bloch
sums of two hybrid orbitals pointing toward one another, or, equivalently, a
Bloch sum of the linear combinations of these orbitals. In the case of &states
the hybrid orbitals are added, and the corresponding linear combinations
I btR) are given by
172 Chapter 2. Electronic structure of ideal crystah

1
I kR)= - 11 h t l
1/2
R)+ I h2Rt + R)], t - 1 , 2 , 3 , 4 . (2.301)

Thc exponential factors eF of the eigenvectors (2.299) and (2.300) were com-
pensated by the et-factors of the Bloch sums. In the case of a-states the
hybrid orbitals are subtracted, and the corresponding linear combinations
I a , R ) are given by

The polar diagrams of these functions are shown in Figure 2.22. For reasons
which will be clarified later, one refers to the orbitals I h t R ) as bondiny and
to the orbitals 1 arR)as anhi-bonding orbitals. With the help of bonding and
anti-bonding orbitals the eigenfunctions of the Hamiltonian matrix (2.296)
with b1' -- 0 may be written in the form

1
I btk) - eik'R I b t R ) , 1 - 1 , 2 , 3 , 4 , (2.303)
aR
(2.304)

In the second step of the diagonalization procedure the Vl-terms are in-
cluded. The Hamilton matrix (2.296) is transformed iuto the basis set of
the previously calculated eigenvectors 1 btk) and 1 atk). This results in the
(8 x 8)-matrix

(2.305)

with the (4 x 4) -matrices


2.6. Tight binding approximation 173

{2.307)

as block elements. The expression for If, follows from that or H & if in the
latter q, is replaced by E,. The structure factors g t p and &) in (2.306) and
(2.307) are defined as follows:

The (4 x 4)-matrix H M couples the various bonding states, and H,, couples
the various anti-bop.ding states. The non-diagonal matrix Hab describes the
interaction between the two types of states. It gives rise to corrections to
the eigenvalues of relative order of magnitude I 11 I /2 I V2 I. Considering
the actual values of V1 and V2. these corrections are rather small. This
suggests treating them as perturbations. The zero-th approximation, i.e. the
complete neglect of the interaction between bonding and anti-bonding states,
is referred to as the b o d ovbital approximation. Within this approximation
the valence and conduction bands follow from separate eigenvalue equations,
the former by diagonalizing the matrix Hbbq the latter by diagonalizing Hw.
To calculate the total energy of the crystal one needs the total energy
of all valence electrons. T h e latter may be calculated by means of formnla
(2.54) which expresses the total energy of an interacting electron system by
means of its one-particle energies. One has

Er$&f= 2 E:(k) -Ecd (2.309)


k z

where E d means the Coulomb energy of the interacting valence electrons


which is counted twice in summing upon all band states. The factor 2
accounts for the two spin states. Within the bond-orbital approximation,
the i-sum in (2.309) may be carried out in closed form, with the result

(2.310)

To pro>*ethis relation, we write the eigenvalues @(k) in (2.310) as diag-


onal elements of H M between eigenstates, and note that within the bond
orbital approximation the eigenstates for a given wavevector k are linear
174 Chapter 2. Electronic structure of ideal crystals

combinations of the bonding states I bik) only. In this matter, the linear
combinations are generated by the unitary transformation which diagonal-
izes the Hermitian matrix H a of (2.306). If one sums the eigenvalues @(k)
over all i , and takes advantage of unitarity, then the diagonal elements of
Hw with respect to the eigenvectors become the diagonal elements of this
matrix with respect to the bonding states, i.e. q,, as is stated in relation
(2.310). One may also prove this relation in another way, using Vieta's the-
orem, which states that the sum of all zeros of a polynomial of degree n
equals the negative of the coefficient of the (n - 1)- th degree term. In
the case of the characteristic polynomial of a matrix, this coefficient repre
sents the negative of the sum of all diagonal elements, here, therefore, -4q,
confirming the validity of equation (2.310).

Total energy and covalent bonding

We proceed on the assumption that the atoms are arranged, as above, in


the form of a diamond type crystal. The distance d between the nearest
neighbor atoms will now be chosen, however, to have a value different from
the do-value of the actual crystal. The total energy of this fictional diamond
type crystal represents a function EEfAt"l(d) of d. We will demonstrate that
E Z Z t a Z ( d )reaches its absolute minimum at the finite distance d = do. This
result constitutes a theoretical proof that the atoms bind themselves into
the form of a crystal, and that the nearest neighbor distance will have the
experimentally observed value. This does not prove the correctness of the di-
amond structure of the crystal, for that was assumed a priori. To verify this,
one must also show that no other crystal structure can yield a lower total
energy minimum. We will not address this question, but rely on experience,
which indicates that elements of the fourth group of the periodic table, under
normal conditions, crystallize into the diamond structure. Our considera-
tions here have the sole purpose of understanding why, in this structure, the
total energy reaches a minimum at a finite distance d = do. In other words,
we want to understand why chemical bonding should occur at all between
atoms of group IV.
The total energy E;f$a'(d) of the crystal is composed of the energy of
the valence electron system in the field of the atomic cores, as well as the
energy of the atomic cores and their mutual electrostatic interaction energy.
For the total energy EE$;f'(d)of the valence electrons of a crystal with G3
unit cells, one obtains from the relations (2.309), (2.310) and (2.298) the
value

(2.311)

To get the total energy of the crystal, the energy of the atomic cores must be
2.6. Tight binding approximation 175

added to the energy value of equation (2.311). In doing so, one may again
use the fact that the core states of the crystal do not differ from those of the
free atoms. This means that only the mutual electrostatic interaction of the
cores results in a structuredependent energy contribution, while the internal
core energies sum to a constant Eo. The core-core interaction energy has,
approximately, the same value as the electron-electron interaction energy
E d between the valence electrons on different atoms. This is true because
the valence electron charge of an atom equals its core charge for the crystals
considered here. The corecore interaction energy approximately cancels,
therefore, against the negative Coulomb energy -E,1 of valence electrons
in expression (2.311). Finally, the total energy of the crystal is given by

(2.312)

The d-dependence of this energy is due to the fact that both Eh and Vz
depend on d - the hybrid energies Eh have d-dependence as they are defined
by the diagonal matrix elements of the Hamiltonian H between Lowdin
orbitals which contain overlap integrals between s- and p - orbitals of adjacent
atoms, and V2 because this quantity is the matrix element of H between
hybrid orbitals at nearest neighbor atoms. V2 decreases with decreasing
distance d, corresponding to an attractive force between the atoms. The
hybrid energy q, increases as d decreases, corresponding to a repulsive force.
For large d, the attraction dominates over the repulsion, and for small d, the
repulsion dominates over the attraction. Overall, the total energy E z l b l ( d )
of the crystal varies with d as shown in Figure 2.24 schematically. At the
equilibrium distance do, it takes its absolute minimum value. This means
that the initially free atoms will not remain free but form a diamond type
crystal with nearest neighbor distance do. They experience what is called
covalent chemical bonding.
In order to provide a better physical understanding of the nature of
covalent chemical bonding, we compare the total energy E g t p l ( d ) of the
crystal with the total energy E f $ T of 2G3 free atoms. For the elements
of the fourth group of the periodic table with their two electrons in atomic
,-levels and two in atomic cp-levels, one has

(2.313)

where Eo, again, accounts for the energy of the atomic cores. The negative
difference of the two energies (2.312) and (2.313) represents the cohesion
eriergy of the crystal. It is given by the expression
176 Chapter 2. Elecbronic structure of ideal crystals

Figure 2.24: Dependence of


energy difference Eza?''' ( d ) crystal atom
-E::+Ts on f,he inter-atomic Etotol -Eta.ral

distance d (schematically).

Formally, the occurrence of a positive cohesion energy is due to the fact


that the matrix element V2 o l H bctween hybrid oxbitals at adjacent atoms
pointing toward one another is negative, and that I 4v2 I exceeds the energy
differericc (cP - cs). The latter difference may he understood as the energy
increase of an atom if one of itb two .+electrons i s lifted into a p-state or,
equivalently, if its four valence electrons a x put into four sp'-kybrid orbitals
rathe1 than into two s- and two p-orbitals. One calls this population the pro-
moted configuration of the atom. In sp'--hybrid states, the electrons of adja-
cent atoms are capable of pronounced interference. This can be constructive
or destructive, depending on whether bonding or anti-bonding states are con-
sidered. In the casc of constructive interference, the probability amplitude
becomes relatively large in the region between the two atoms and the two
electrons of the interfering sp3-4ybrjds undergo a delocalization (see Figure
2.22). In this process, the potential energy of the Coulomb interaction of
the two electrons among themselves and with the atomic cores remains al-
most unchanged. However, their kinetic energy decreases considerably. This
may be imderstood in terms of the Heisenberg TTncertainty Principle which
tells us that a weaker localization, i.e. a larger positional uncertainty, cor-
responds to a smaller momentum uncertainty and, therefore, to a smaller
kinetic ene~gy. Altogether, the energy of the two electrons decreases, b e
cause of constructive interference, in a bonding state. The energy gain per
atom amounts to 4 I Vz 1. If it exceeds the energy necessary for promoting
an atom into its sp'-state, i.e. if the condition 4 I V2 I> ( e p - F ~ )holds, it is
energetically favorable for covalent chemical bonding to occur. As we have
seen, quanturti rtieclianical phenonienology is essential in the interpretation
of this behavior. Unlike the bonding of electrically diflerent charged ions,
covalent bonding between neutral atoms Lannot bc understood in terms of
classical physics.
The c:ondition necessary for thc occiirrcnce of bonding eigenstates able to
host all vrtlprrcp electrons is the ordering of the newest, neigIi1,ors of an atom
on the comers of a tetrahedron, i.e. the diamond structure of the crystal.
In this way thc above consideration also justifies focming on the tetrtlhedral
crystal structure of diamond type crystals, which was merely assumed at
the outset. The atomic structure follows, so to speak, from the electronic
structure.

Ionic bonding

The ionic contributions to chemical bonding will now be calculated for mat,+
rials having the zincblende structure. A s is well-known, a scries of 111-V and
11-VI compound semiconductors form crystals of this type. For the Ilamil-
tonian matrix (2.305), the transit,ion from the diamond to the zincblendc
structure means that ch in the upper left (4 x 4)-block has to be replaced
by the hybrid energy EL of the 1-atom? and in the lower (4 x 4 )-block by
the hybrid energy e i of the 2-atom. With this replacement, the bonding and
anti-bonding energy levels become

where !+,
=c i - ~8.Thr energy separation betwren the two levels is larger
than that of diamond type crystals. This results in an enlargement of the
energy gap betwcm the valence and conduclion hands. The bonding and
anti-bonding oibilals arp given by the expressions

where we set a p - Vs/dV; + bp. T h e f d o r s {1/2)(1- a p ) and (1/2)(1+ctp)


in ('2.316) and (2.317) represent the probabiMes of finding an elrctron in
the bonding state at atom 1 or 2, respectively. One calls aP the polarity uj
bonding orbitals or simply the polardty of bonding. If f f f is deeper than c i , !
then I!?, and also up,art' positive. The electron prpferpntially stays at atom 2.
In this way the polarity of bonding orbitals is such that, in the ground state,
whcre the electrons occupy only boxrding orbitals, the previously electrically
neutral atoms heconie charged. Atom I becomes the positive cation, and
atom 2 is the negative anion. The charge of the cation is given by c Z *
with Z' = (Z, - 4 t $ap),where 21 is the number of valence electrons at
178 Chapter 2. Electronic structure of ideal crystals

thc free 1-atom. The anion charge is -e(Zz - 4 - 4 a p ) = - e Z * , i.e. the


unit cell is neutral. Owing to this redistribution of electron charge, the
electron-electron interaction energy to be subtracted from the sum of one-
particle energies, because of double counting, takes a different, value. It is,
therefore, no longer completely compensated by the electrostatic interaction
energy between atomic cores. This leads to an additional contribution to
the total energy of the crystal which may be interpreted as the electrostatic
interaction cnergy between anions and cations. Onc calls it the Madelony
energy E M & - The general expression for E M a d iu

where the sum extends over a periodicity region. With EMad, the total
energy of the crystal is

The Madelung energy is negative, i.e. it strengthens chemical bonding. Since


the bonding is then pdrtially due l o attractive forces between ions, OTW refers
to it as partially zonzc bondzng. The absolute value of the Madelung energy
is, on the one hand, proportional to the number GT of unit cells, and on the
other hand, inversely proportional to the distance d between two adjacent
ions. One therefore sets

(2.320)
with cy as the so-called Madelung constant. The latter depends on crys-
tal structure and can easily be calculated numerically. In Table 2.11, the
n-values are listed for crystal structures which are observed in materials
composed of group IV elements as well as 111-V, 11-VI and I-VII compounds.
The value for the wurtzite structiirp in Table 2.11 corresponds to the ideal
tetrahedral case with an equivalent cubic lattice constant &a (see Chap-
ter 1). The contribution of the Madelung energy to the total cnergy of a
given compound will be larger for larger effwtive charge number Z* of the
compound. This results in a tendency of compounds with larger 8* valiirs
to crystallize in structures with Madelung constants larger than that of the
zincblende structure. Therefore, in passing from the 111-V through the II-
VI to the I-VII compounds, one observes a transition from the zincblende
structure through the wurtzite to the rocksalt and cesium chloride struc-
tures. The cesium chloride structure follows from the rocksalt structure by
replacing the two facecentered cubic sublattices by two primitive cubic sub-
lattices, shifted in the same way with respect to each other as in the rocksalt
2.7. k . p -method 179

Table 2.11: Madelung constants for several crystal structures.

Zincblende 1.6381
Uurtzite 1.6410
Rocksalt 1.7476

structure, i.e. by ( a / 2 , a / 2 , a/2). With growing polarity of the bonding, the


energy gap becomes larger, as mentioned above. This explains the transition
from the semiconducting properties of the group IV crystals to the insulating
nature of the I-VII compound crystals.
In the case of the I-VII compounds, the absolute values of V3 are so
large in comparison with V2 that the bonding polarity op is approximately
unity. This implies that almost all valence electrons of the compound stay
at the anion. Then the crystal consists of positive ions of the group I atoms,
which have lost all their valence electrons, and negative ions of the group
VII atoms whose valence shells are completely filled. One refers to such
crystals as tonac crystals. In this case, the energy gain due to the transfer
of electrons from cations to anions, which represents an essentia1 part of
the bonding energy and forms the driving force for the formation of ions,
no longer depends on the crystal structure. This structure is determined
by the hladelung energy only. Therefore, ionic crystals exhibit structures
with particularly large hladelung energies, i.e. rocksalt and cesium chloride
structures.

2.7 k .p -method

2.7.1 Fundamentals

Luttinger-Kohn functions

The k. p-method rests on a particular property of the BIoch type eigen-


functions pV,+(x)of the crystal Hamiltonian H . As we know these functions
(which will be denoted below by (xluk) instead of q y k ( x ) ) are the product of
an exponential factor exp(ik.x) and the latticeperiodic BIoch factor uvk(x).
If one replaces the wavevector k in uYk(x) by a constant ko, while retaining
k in the exponential factor, then the resulting functions
1x0

[2.321)

are no longer eigenfunctions of H of course, but they do form a complete


~

orthonarmalizPc1 basis set in Hilbert space, as wcll as the Bloch functions.


whence

(vkko(vkko)- 6 v l u 6 k ~ k , (2 322)

~ ( x l v k k o ) ( v k k o l x--) b(x - x). (2.323)


uk
Thp vdidily of these relations f d o w b directly from the rorripletenyss and
orthonormality of the Bloch functions. The (xlvkko) are referred tu as
LuftmgGgP7- K u h fiLa~t7onu.T h y arp determined by the Blorh factors u V k ( x )
for the special wavevector ko in contrast to the Bloch functiom which
require full knowledge u l u,k(x) lor all wavevectors k. The k . p-method
takw advantage of this properky of the Luttinga-Kohn functions.
In this method, one represents the Srhriidinger quation for R crystal
electron in terms of the complete orthonormaIized set of these functions.
The rpsulting matrix elcments of H can be expressed, as we will s e e later,
by the matrur elements of H between the Bloch factors Uyk(X) for k = ko.
These elements arc, of course, just as little known as the Bloch factors them-
selves. However, one may take them as empirical parameters. If one does so
and inserts values for the parameters. then the Hamiltonian matrix in the
Luttinger-Kohn basis is completely determined. Uiagonalizing this matrix
yields the eigenvalues and eigenfunctions of the crystal Hamiltonian H for
all valiies of k. Tllis means that the k p-method allows 011rto calrulate,
from the Bloch matrix elements at only one point kol the eigenr-alues and
cigenfunctions over the entirc first 32,i.e. to extrapolate from thc particu-
lar point ko t o t h e entire first BZ. ()ten one is only interested in solutions
in the vicinity of a critical point k,e.g. in the vicinity of the valence band
ninxinium or the coridurtion band minimimi. Then it is expedient. although
not necessary, to icienliljr ko with k,. If k, hes. for example, at the center
of the 6rst BZ,as often occurs. one has ko = 0. This choice will be used
later. At the outset, ko should still be considered an arbitrary point of the
first HZ.
In order to accomplish the ahove program. we expand thP Bloch functions
(xjuk}with respect to Luttinger Kohn functions (xjpkko). On\y terms with
k = k occur in this expansion because of the lattice translation symmetry
of both functions, whence
2.7. k . p -method 181

With this expansion, the Schriidinger equation (2.178) in the Luttinger-Kohn


representation bmomes

k .p - Hamiltonian
lhematrix elements (pkko I H I pkko) of the llarniltonian bptween T,uttingpr
Kohu functions can he t r a d back to matrix elements (pko I p j pko) ol the
momentum operator p between Bloch functions. if one uses the easily provcn
commutation relation

[p2,eik-x] = eik-x (p2 f 2fik. p + h2k2) , (2.326)


which yields

where we have set

Ti2
E:(k) = E,(ko) + -(k
2m
- ko)2. (2.328)

The matriv on the right-hand side of (2.327) allows for an important rewrit-
ing. If one defines

Hk.p(k) = Ho(k) + -nah(k - ko) . p! (2.329)


with

(PkkolfflPkko) = (PkoIHk.,(k)lPko). (2.33 1)


The latter relation means that the actual Hamiltonian matrix W in the k-
dependent Luttinger-Kohn basis Ipkko) equals the representative matrix of
a fictional k-dependent Harniltonian Hk.p(k) in the k-independent partial
Bloch basis lpkoko) = jpko) for the wavevector k = ko. The k-dependence
of the Luttinger-Kohn basis on the left hand-side of equation (2.327) has
been transferred to the new Hamiltonian Hk.p(k) on the right-hand side.
The SchrGdinger equation (2.325),with this new Hamilt onian, reads
182 Chapter 2. Eiectronic structure of ideal crystals

The components of the eigenvectors Ivk) in (2.332) refer to the Luttinger-


Kohn basis Ipkko), although the operator Hk.p(k)is represented in the Bloch
basis Ipko).
Solution of the Schrodinger equation (2.332) involves the diagonalization
of the matrix (pkolHkE,(k)lp'ko). For k = ko this matrix is automatically
diagonal, by virtue of the fact that Bloch functions Ipko) are eigenfunctions
of the Hamiltonian Hk.p(kO) = Ho. For k # ko, the Ipko) states are no
longer eigenfunctions of Hk.&), so that the matrix (pkolHk,(k)Ip'ko) has
off-diagonal elements with respect to the band indices. Formally. one may
interpret these non-vanishing elements as arising horn an interaction between
different bands. Since this interaction results from the (k - ko) . p-term
in Hk.p(k), one calls it the k . pinteractian. In this, the bands which are
mutually coupled, are not bands in the sense of the eigenvalues of the actual
crystal Hamiltonian H - the latter are uncoupled by definition - they are
fictional bands E:(k) defined by equation (2.328).
As the point k in Hk.,,(k) approaches ko, the k . p-interaction tends to
zero. For k-vectors sufficiently close to ICO, one can treat this interaction
with the help of quantum mechanical perturbation theory. Apart fkom the
square term in (k - ko) already present in E:(k), this entails a power series
expansion of the energy bands E,(k) with respect to (k - ko) about the
point ko. The form of the perturbation theoretical expansion depends on
whet her the unperturbed bands, i.e. the eigenvalues E,(ko), are degenerate
or not. We will first consider the simpler case of non-degenerate bands.

Application to non-degenerate bands. Effective masses

In first, order perturbation theory the eigenvalue E$(k) arising from EE(k)
is given by the relation

(2.333)

and the Bloch function Ivk)' arising from 1.k)' E Ivkko) by the relation

Since the f m t derivatives VkE,(k) of the exact band energies E , ( k ) at ko


depend only on linear expansion terms of E,(k) in k - ko,no approximation
is needed to obtain the relation OkE,(k)lk, = VkEb(k)lb. Considering
2.7. k . p method 183

(2.3331, this exact relation yields V&v(k)lb = ( h / r n ) ( v b l p l v h ) . This


holds the same content as equation (2.193) used above without proof, because
ko may be an arbitrary point of the first B Z . In particular, if ko is a critical
point kc, i.e. if V&,(k)\k,, = 0 holds, then the k .p-correction vanishes in
first order perturbation theory. One must proceed to the second order to get
a non-vanishing contribution from the k p-perturbation. The result reads

where, for brevity, we set (I& I p I pkc) = (v 1 p I p ) . We rewrite this expres-


sion in the form

Generalizing the terminology introduced in section 2.6, we call M L 1 the


effective mass tensor at the critical point k,. For the diagonal elements
mz' of At;' with respect to the principal axis system, one obtains from
(2.337) the relation

This relation connects the effective masses with the matrix elements of the
momentum operator between different bands and with the energy separation
of bands at the critical point. The tendency indicated is that the absolute
values of the effective masses become larger for smaller momentum matrix
elements and larger band separations. One expects small effective masses
for large momentum matrix elements and small band separations. As far as
the band separations are concerned (only for them can one make an easy
estimate), we will later find conhmation of this tendency in all concrete
cases. For pairs of bands which are closer to each other than to all other
bands and, therefore, whose mutual interaction is stronger than that with all
other bands, relation (2.338) allows one to also draw a conclusion about the
signs of the effective masses. According to it, the energetically higher of the
two bands should have a large positive effective mass, and the energetically
lower a mass of the same large absolute value but of negative sign. This
184 Chapter 2. Electronic structure of ideal CFZ.S~~~S

conclusion also proves to be valid in all cases in which the assumptions of


this calculation apply.

Band degeneracy

Critical points are often symmetry centers or lie on symmetry lines, and
at these symmetry points, degeneracy of the energy bands often occurs. If
this happens, one must carry out second order k.pperturbation theory or
degenerate bands. In quantum mechanics, perturbation theory for degener-
ate energy levels is c o m o n l y of first order - the matrix of the perturbing
Hamiltonian operator between the degenerate states has to be diagonalized
(we remind the reader of the nearly free electron approximation in section
2.4). This procedure does not apply here because the perturbation matrix at
critical points vanishes in first order. One must therefore choose a variant of
perturbation theory for degenerate energies which works in second order. To
this end, one constructs the matrix of the perturbation operator not between
degenerate unperturbed eigenstates, as is commonly done. but between the
(also degenerate) eigenstates of first order of perturbation theory. By diago-
nalizing this matrix one obtains the eigenvalues in second order perturbation
theory. These are, in general, no longer degenerate.
An important case in which the k . p-perturbation matrix between the
degenerate unperturbed states vanishes, is the valence band maximum of
semiconductors with diamond structure. This case will now be investigated.
In doing so, w e initially neglect the spin-oIbit interaction. This approxima-
tion is valid for semiconductor materials composed of light elements only,
including, for example, Si. For other materials this procedure serves as a
zero-order approximation which can be used to proceed further (as we will
do below).

2.7.2 Valence bands of diamond structure semiconductors


without spin-orbit interaction
As we know from section 2.3, the valence band maximum of diamond type
semiconductors is located at the center r' of the first B Z . Therefore, we set
& = 0. The maximum is 3-fold degenerate. We denote the three pertinent
Bloch functions by IvmO), where m can assume the values 2,g,2. According
to section 2.6, these eigenfunctions belong to the irreducible represent ation
l?b5 of the cubic group oh. As indicated in Appendix -4. a basis of this rep-
resentation is formed by the products yz, zx,z y of the components 2,y, z of
position vector x. Therefore, with regard to their transformation properties
under the action of elements of o h , we may identify IvzO) with yz, IvyO) with
zz, and IvzO) with zy. The vector components 2,y, z of the position vector
itself transform in accordance with the irreducible representation I'15 of Oh.
2.7. k . p -method 185

For the subgroup T d of oh the two representations F15 and r h 5 coincide. For
semiconductors having the zincblende structure, the three degenerate states
( v I O ) , IuyO), IuzO) of the valence band maximum may therefore be associ-
ated with I,y, z insofar as their transformation behavior is concerned. In
the case of the diamond structure, 2,y, z are merely a short hand notation.
The vanishing of the matrix (vmOlplum'0) of the momentum operator
between valence band states at I?, anticipated above, may easily be demon-
strated using the pertinent criterion for such vanishing given in Appendix A:
The operator p transforms according to the irreducible representation I'15
of oh. The matrix (umOlplvm'0) therefore belongs to the reducible repre-
sentation + + +
x r 1 5 x rl,, = I?;, x (rh ri2 I'15 r 2 5 ) , wherein the identity
representation does not occur. According to Appendix A this means that
the matrix (umOlplvm'0) must vanish. One can also obtain this result by
means of inversion symmetry alone. We have chosen the somewhat more
troublesome method of proof because it may also be applied in other, less
obvious cases, as we will see immediately below.

k . p -perturbation theory t o second order with degeneracy

In order to apply degenerate second order perturbation theory, the solu-


tions of Schrodingers equation (2.332) are needed to first order in the k . p-
perturbation. For the orthonormalieed Rloch valence band eigenstates (vmk)l
one finds

(2.339)

where we set E,(O) = for brevity, and the degenerate valence band energy
E J O ) is denoted by Ev. The third term in (2.339) guarantees the normal-
ization. (:onsidering the sum on p , tlir value 11 = 71m does not need to be
specifically excluded because the matrix elements ( p I p I vm') for p = 117n
vanish anyway. Expressions of tlic form (2.339) also hold for the approxi-
mate Bloch functions Ipk)' of the remaining bands p with p f urn, but we
omit an explicit presentation of them here. The states Ivmk)' and Ipk)'
with p # t m will now be used as a basis set to represent the IIamiltonian
H . The resulting matrix is 'almost' diagonal, because the basis functions
are 'almost' eigenfunctions. In particular, the submatrix of the three velence
bands is coupled to the remainder of the matrix only by elements of second
order in the k . p-perturbation. These elements give rise to corrections of the
valence band energies which are only of third order and can be neglected.
186 Chapter 2. Electronic structure of ideal crystals

In second order perturbation theory, the valence band energies Ev and For-
responding Rloch states I&,) therefore follow from an rigenvalue equation
which is decoupled from the remaining bands, namely

'(vmklH!vm'k)l l(wn'kl&:y)= E ' , '(vrnklE,). (2.340)


rn,

The initial occurrence of interaction betwren the valence band and the r e
maining bands is incorporated in the matrix (u>mk I H I wn'k) in first order
perturbation theory.

Harniltonian m a t r i x

The (3 x 3)-Hamiltonian matrix '(vmklHlvm'k)' of equation (2.340) can be


obtained by means of expression (2.339) for the perturbed states Ivrnk)'. A
short calculation yields

where

is a fourth-rank tensor. Since the states I p ) = 1 P O } are eigenfunctions of


H with eigenvalue EEl= E,(O), one may write (2.342) in the more compact
form

With respect to the indices a r P lthe tensor D z i , is symmetric, and with


respect to the indices m , m' it is IIermitiaa From equation ('2.345) one can
see that D Z k , transforms under symmetry operations of the cubic group
Oh according to the $-fold product representation [rb5 x ri5]a
x [1'1~ x l ? 1 ~ ] ,
where the index s denotes the fiyrnmetrical part of the product. According
to Appendix A, L?zk, then contains as many independent elements as the
number of times the identity representation occurs in the product [I?& x
riE;lrn = [I& x r&lS
x rl5Is
x [I'Is x I115Is. Using Appendix A , one finds [I'15 =
+ +
r1 r12 ra,, which yields [rk5x r& x [rI5x rl& = 3r1 rz 4r12 + + +
3c,-k 5T'k5 The tensor D z L , therefore has three independent components.
one can show that these correspond t o the three types of non-vanishing
2.7. k.p -method 187

matrix elements D z z , Ll;& and D Z . We introduce the abbreviations L -


D g , hl = D$$, and N = D$ f D g . The elements L, M and can be
calculated if the Bloch factors are known. In the absence of this informalion,
however, we consider L , M and N to IIPempirical parameters (as indicated
at the outset) and use their connection with the Bloch factors only to identify
some general properti=, such as the fact that they can be chosen real. Since
all remaining matrix elements Dmm, Q B vanish, the Hamiltmian matrix of the
valence band has the form

Method of invariants

The Hamiltonian matrix (2.344) can also be derived in a somewhat different


way, whirh leads to the goal more quickly, but is formally more dernand-
ing. One uses the fact that the Hamiltonian matrix l(vmklHlum'k)l can
be represented as a linear combination of the 9 matrices of a basis in the
product space 1 vmO)(iirn'O 1 which transforms according to the repreaenta-
tion [I& x r:,] of the point group oh. This representation is reducible. By
decomposing it into i t s irmliicible parts, m e obtains a basis which consists
of subbases, each of which belongs to a particular irreducible representation
of Oh. Such a matrix basis can easily be constructed by means of the 3-
dimrnsionel angular niomenturn matrices Iz,Iv,Iz (considered in Appendix
A ) and their products, since it is known how these matrices transform,
namely according to the pseudovector representation I':s. In the product
spacc k,f-fi of the components of the vwtor k, one proceeds in a similar way.
One determimes a basis from subbases which transform according to the ir-
reducible parts of the representation [I'15 x r15Iy.The Hamiltonian matrix
reprewrits an element in the product space of Ihe two spaces which is invari-
ant under transformations of the point group Oh. Such invariant elements of
the product space can be produced by forming scalar pruducts of subbases
of the two spaces whi& transform according to the same irreducible repre
sentation. As seen in Appendix A, the corresponding scalar products belong
to the- identity representation, that is to say, they are invariant.
To find the most general Hamiltonian matrix compatible with the sym-
metry oh, one has to determine all invariants of the product space. I. one
then multiplies each by a real scalar factor and sums them all, one obtains
188 Chapter 2. Electronic structure of ideal crystals

the most general invariant of the product space and thus the most general
Hamiltonian matrix compatible with Oh symmetry. This process is called
the method of invariants. It is applicable to arbitrary symmetry groups and
degrees of degeneracy, and it quickly leads to the goal if one considers spin
and spin-orbit interaction. It also allows one to determine the matrices for
perturbing Hamiltonians other than that of the k . p-interaction, such as
the interaction between the angular momentum of Bloch electrons and an
external magnetic field (see section 3.9) or the interaction with mechanical
strain. In this book, we will only use the method of invariants occasionally.
A comprehensive outline of the method with several applications i s given by
Bir and Pikus (1974).

Valence b a n d s t r u c t u r e

The eigenvalues of the matrix (2.344) form three valence bands E,l(k),
E,n(k), E,3(k). For the three symmetric k-directions [loo], [lll]and [110]
the dispersion curves are determined by simple analytical expressions as fol-
lows:

E,lp(k) = M k 2 , E,3(k) = L k 2 , (2.345)

1 1
E,lp(k) = - [ L
3
+ 2M - N]k2, E,3(k) = - [ L
3
+ 2M + 2 N ] k 2 , (2.346)

E,l(k) = M k 2 , (2.347)
1
E,2(k) = - [ L
2
+ M + N]k2, E,3(k) = - [ L
2
1
+M - N]k2. (2.348)

Along the two directions [loo] and [lll], the valence band, being triply
degenerate at r, splits into two bands, one 2-fold degenerate and one non-
degenerate (see Figure 2.25). In the [llO]-direction and also for all more
asymmetric k-vectors, no degeneracy remains. This indicates a %fold split-
ting of the valence band for such k. All bands are parabolic, but evidently,
in general, not isotropic. One speaks of a warping of energy bands. In the
case of Si, one has L = -5.64, M = -3.60, N = -8.68 in units of ( h 2 / 2 m ) .
Using these values, the two degenerate bands E,l/a(k) of equations (2.345)
2.7. k . p method 189

Figure 2.25: Valence band dispersion for diamond type semiconductors in the
vicinity of the I'i5-maximum for different k-directions.

or (2.346) have smaller curvatures than the third band E,s(k) in these equa-
tions. Thus the first two bands correspond to the heavy holes and the third
band to the light holes of Si. Isotropy exists only if N = 0 and L = M .
Then, there also is no longer any distinction between light and heavy holes.
Conversely, anisotropy grows stronger as the difference between the masses
of the two types of holes becomes larger.
The results discussed above were obtained without consideration o l spin-
orbit interaction. However, for most of the diamond and zincblende type
semiconductors, the valence band structure is significantly influenced by this
interaction (in the case of Si it is small, hut often not negligible). We now
proceed to consider the effects of spin-orbit interaction.

2.7.3 Lut t inger-K ohn model

Including of spin-orbit interaction

To include spin-orbit interaction the electrons must be treated as particles


with spin. The electron states then depend on both the position coordinate
x, and on the spin-coordinate 3, and the quantum numbers of the electron
states contain provision for both the motion in coordinate and spin state u
describing the spin motion. A set of basis functions for representation of
the electron states in coordinatespin space may be formed lrom the Bloch
functions (xlpk) or the Luttingw-Kohn functions (xlpkO) without spin, aug-
mented by multiplication with the eigenfunctions ( s l u ) of the z-component
of the spin operator. Here u means the spin quantum number which can
have the two values u -r (spin up) and cr =I (spin down). We denote the
product functions by (sxlpk) or (sx(pkO),respectively. Then, by definition,
190 Chapter 2. Electronic structure of ideal crystah

we have

The total Hamiltonian of the system is obtained by adding the spin-orbit


interaction H, to the Hamiltonian H in its absence, where Hgo is given by
equation (2.56) as

(2.3501

One has to be aware that the (sxlvuk) are eigenfunctions of H , but not
of H +H,. Correspondingly, the (xlvuko) signify the spin-dependent
Luttinger-Kohn functions of H , but not of fi + f l W .T h e matrix repre
senlation of the Schrtdinger equation with respect to the spin-dependent
Luttinger-Kahn basis reads

x(prkOtH + H,, IpvkO)(pukOIE l , ) - E , (pu kO IE,) . (2.351)


pW

In calcuhting the matrix of H t H,, of (2.351)) an additional k-dependent


term appears in comparison with the spin-less case, as a consequence of the
fact that H , contains the momentum operator. This term has the same
form as the (A/rn)k. p-term arising from H . except that the p-operator is
replaced by the operator p+(1/4rnc2)[5 x OV(x)]. The additional term can
be taken into account by replacing the operator Ht.p of equation (2.329) (for
ko = 0) by the operator
Tl
Hk.x = H o ( k ) + -k
m
. a, (2.352)

with
1
ii = p + ---[a
4m c=
x VV(X)]. (2.353)
With this the Luttinger-Kohn representation of H + H,, becomes

(pukOlH + H80/pakO) = ( ~ u O I H+~ Hso{puO),


.~ (2.354)
and the Schrodinger equation (2.351)) takes the form

C ( ~ ( T O I H+ ~HsoIpuO)(pukOIEu)
.~ = E,(pnkOlE,). (2.355)
PU

Up to this point, we have kept the discussion general. Now we wish t o ex-
plore the particular consequences of spin-orbit interaction for the previously
2.7. k p method 191

considered valence band states. To this end, we need the matrix ele-
ments (vrnuO~Hso~wmuO) of H,, between the spin-dependent Bloch states
Ivm 0 . 0 )namely,
~

F,
( u ~ u O I H , , I ~ ~ U =O ~) ( v v ~ [VV(X)
O I x p] IumO) x (aldld). (2.356)
4m c

To evaluate the matrix element (vmOl[OV(x) x p]]vmOj in coordinate space


we make use of crystal symmetry, as was done before, in the calculation of
the matrix elements of the momentum operator p. The operator [VV(x) x pi
is a pseudovector and transforms according to the irreducible representation
ri5of oh. The entire third-rank tensor (vrnOI[VV(x) x p]lvrnO) therefore
belongs to the reducible representation Ib5 x Ti, x P2,= rl,, x (r2 r12+ +
Ti5+I7L5), in which the identity representation occurs exactly once. The ten-
sor (vmO ! [VV(x) x p] I vmO) consequently contains one independent con-
stant. This constant coincides with the matrix elements (vyOl[VV(x) x
pIz I v z O j = (vzO I [VV(xj x plXI vyO) = (vzOI [VV(x) x p]ylvzO), as well as
(~.Ol[V(X)XPl*l v y 0 ) = (WYOI [ w x ~ x P l . / ~ =
~ (oV Z) O l PV(XjXp1ytvrO).
where (vxOl[VV(x) x plz I u y O ) = -(uyO I [VV(x) x p],lvzO) holds. Because
the Bloch factors are real? these elements are pure imaginary. We denote
the value of (vyOl[VV(x) x p],jvzO) by (4m2c2/h)(i/3)A, i.e. we set

h A
,(uyO~[VV(x)x plrlvzO) = i - . (2.357)
4m c 3
Below, we will see that the constant A is the energy splitting of the valence
band at J? due to spin-orbit interaction. Applying equation (2.357) and the
explicit form of the spin matrices given in (2.57),the matrix (avmOlH,I
aw m 0j becomes

0 - i 0 0 0 1
i 0 0 0 0 - 4
A 0 0 0 - 1 i 0
(vmaOp,,Ivm.uO) =- (2.358)
3 0 0 - 1 o i 0
0 0 - i - i o 0
1 i 0 0 0 0

Here the rows end columns are associated with the basis functions in the
sequence Ivz T O j = Iz t), Ivy 7 0 ) = ly tj, . . ., It)z 1 0 ) 3 1s 1). Spin-
orbit interaction couples orbital states and spin states to each other. At
k = 0 the expressions E:(k) and k ii are both zero. The eigenenergies and
eigenfunctions of the total Hamiltonian Hk.a +
H , are therefore also those
192 Chapter 2. Electsonic structure of ideal crystals

of H,, alone, so thsl the Schrfidinger equation (2.355) at k - 0 becomes

c
u'm'
= E(wamO(E).
(vmaO~H,,lvm'a'O)(vm'a'O~E) (2.359)

Eigenfunctions at r. Angular momcntum basis.

The matrix (2.355) has a 4-fold degenerate eigenvalue


1
Ev1/2/3/4 = p (2.360)
with the four corresponding eigenstates, respectively,

1
IE ) T ( 1 ,i, o,o, O,O), IEv2) -
v1 -
~

Jz
1 i
IEv3) = z(l, - i , O , 0 , 0 , 2 ) , IEv4) = -(O, O,O, 1, - i , O ) , (2.361)
Jz
and a 2-fold degenerate eigerivslue
2
Ev5/6 - --A (2.3 62)
3
with the two corresponding eigenstates
1 i
IEv5) - -(o, 0, 1, 1, i, o), 1E ) - ----(--I, i, o,o, 0, 1). (2.363)
fi w6 - d3

The components of the eigenvectors given by (2.361) to (2.363) refer l o the


basis functions IvmaO). If we abbreviate these by Ima) IwmaO), using
Iz T), J y I), 12 I), Iz J), ly I ) , Iz i), the cigcnvcctors take the form

The eigenvectors lEvt),i - 1,2, . . . , 6 of (2.361) have a simple meaning. The


lEvl),IEvz),IEv3), IEv4) are basis functions of the irreducible representation
of the cubic group o h . Acrording to Appendix A, these representations
2.7. k p -method 193

emerge from the representation 213 of the full rotation group if, D Dis~taken
as a representation of the subgroup oh with +1 for inversion. It has also been
shown that the basis functions of this representation are the simultaneous
eigenfunctions of the angular-momentum-squared ' 5 or the eigenvalue j ( J +
1) (in units R 2 ), and of the z-component J z of J for the eigenvalues m j =
- -
L and -$
23 , 12 ., - 1 . =- 4.
One therefore denotes the first four eigenfunctions
of (2.364) by

33 1 31 i
I--)
22
= -1rz
4 + iy t), I--)
22
= -"-2Iz
fi t) + la: + iY 1 1 1 7 (2.365)

3T 1 33 i
I--)
22
= -[&
I. - iy T) +2/2 111%I--)
22
= 1.-
Jz -iY I),
The lEvs),IE,s) are basis functions of the irreducible representation I'y of
o h . These representations do not arise from any representation V3 of the
full orthogonal group. in particular not from a representation for J = 4
(this happens with rt
in the case of Oh or r:) in the case of T d . but the
expectation values of J2 and J z are the same as those in the Dl basis.
Therefore one also uses the angular momentum notation or the last two
eigenfunctiona of (2.3641, i.e. one sets

Each of the dgenvectors (2.366 and (2.366) is determined only up to a phase


factor, which is chosen heie such that the states with negative total angular
momentum, ?I):, $I,): 1 follow, respectively: from the states with positive
total angular mornenturn, I$$), ,);I ) ; ; 1 by means of time reversal, i e . by
forming the comphx conjugate of the original eigenvector and subsequently
multiplying it by rP.
We refer to the functions Ijmj) of (2.365) and (2.366) henceforth as the
angular momentum basis. According to (2.360) and (2.3621, eigenstates hav-
ing the same eiggmvalue of the the angular momentum squared, J2,also have
the same energy eigenvalue, while the rncrgy eigenvalues differ if states with
different eigcnvalues of J' are consitirrrd. The valence band: being B-fold
degenerate at the r-point if spin is not taken into account, therefore splits
into two bands, one with j : and one with j = 3, 4,
if the spin-orbit in-
teraction is considcrccl. That such a .ynin,-o)rhitqditting must occiir? one can
recognize just by means of a grnnp theoretical analysis of the problem. The
r
six valence hand states at transform in accordance with the &dimensional
representation Dl x , ;'I of Oh. Tlicse representations are reducible, accord-
ing to Appendix A , as Di x
T
= Y$ + I';. The size of the splitting is given
194 Chapter 2. Electronic structure of ideal crystaJs

by the constant A, which determines the strength of the spin-orbit interac-


tion. Therefore A is called spin-orbit splitting energy. One may interpret A
as the difference of the spin-orbit interaction energy between the states with
j = 33 and those with j = i. As one should expect, the states with larger
angular momentum lie energetically above the states with smalIer angular
momentum. States with different m j , i.e. with different projections Jz of
total angular momentum on the z-axis, but the same J2,have the same spin-
orbit interaction energies. Therefore the degeneracy of these states remains.

Valence band structure off r

The above statements refer to valence band states at the center r of the first
B Z , where the k p-interaction vanishes. Off r,this interaction is no longer
zero and must be taken into account in addition to the spin-orbit interaction.
We have seen how this can be done approximately in the preceding section,
without consideration of spin. The method used there indicates the following
procedure in the presence of spin and spin-orbit interaction: One determines
the functions Ipok)' which diagonalhe the operator H k q of (2.352) in first
order perturbation theory. In analogy to equation (2.339), one finds for the
valence band states Ivmcrk)' the expression

where we use the same abbreviations as in (2.339). Analogous relations hold


for the states of the other bands. The functions Ipgk)' form a complete
orthonormal set in t a m s of which the Hamiltonian H may be represented.
The submatrix with respect to the valence band states (vmmk)' is decou-
pled from the remainder of the matrix in second order perturbation theory.
Since H,, also only couples the valence band states among themselves, but
not to states from other bands, the Schrdinger equation (2.355) in this
representation reads

The spin-dependent term of k .7i in the eigenstates Ivmcrk)' of (2.367) d e


scribes the change of the k .p-interaction due to spin-orbit interaction. '5'ince
t,he two interactious are supposed to be weak, this change i s second order
small. It will be omitted below. Then we have, approximately,
2.7. k p -nrt.th<>d
9 195

(vmakIH,,lv7nak) (vmaOlH,qolvmaO).
Now we use the fact that H,, is diagonal in the angular momentum basis
Ijmj) of (2.365) and (2.366). It is clear that this basis follows from Ivma0) E
Imu) by a unitary transformation

~ j m j ) C U r r m j m j Ima). (2.370)
ma
The corresponding unitary transformation matrix UmjmJ can be readily
obtained from the rclations (2.365) and (2.366). One has

(2.371)

If one applies this transformation to the Schrodinger equation (2.355), then


the matrix (vmaOlH,,(vmaO) takes the diagonal form

o $ o o 0 0
o o + o 0 0
(hjlff,oljm;) = (2.372)
o o o g 0 0
0 0 0 0 - 9 0
196 Chapter 2. Electronic structure of ideal crystds

The sum of the two matrices (2.372) and (2.373) is the new Hamiltonian
matrix. It has the same eigenvalues as the original matrix, even though its
form deviates from that of the original. The difference in form is, above all,
that the new matrix is already diagonal at k = 0. Kon-diagonal elements
occur for k # 0. Among them, the elements between basis vectors ] j m 3 )and
[ j ' r n $ ) with m3 # m i , , but = 3' play a different role than the ones with
.f j'. While the influence of the 2-diagonal elements on the eigenvalues is
independent of the size of the spin-orbit splitting A, it does depend on it for
the j off-diagonal elements. The magnitude of the latter can be estimated
as the larger of the two terms Nlki2 or IL - Mllkl'. If one assumes that
l L f a ~ ( ~ VIL, - AIl}lk12 << A holds, then a perturbation theoretical treatment
is possible. It yields an energy correction of order of magnitude [ M a r { N ,IL-
M1}]21k14/A. Under the assumptions made, this is small compared to A.
That means that the 3-off-diagonal elements of the transformed Hamiltonian
matrix can be neglected if the k-vectors are sufficiently close to r. We will
assume below that this is the case, although the Luttinger-Kohn model also
covers the general case of a (6 x 6 ) -Hamiltonian matrix. Neglecting ]-off-
diagonal elements the Hamltonian decomposes into two blocks, one (4 x 4)
-block conesponding to the basis vectors of the representation I'i,and one
(2 x 2)-block for the basis vectors of the representation I't.
The rsf-matrix
reads

Here, therows a n d c o l u r r ~ u a r e o r d e r e d i n t , ~ ~,):;I~ ~ u,];I~ c e,)::I ,)::I


9
thc 1 multiplying A/3 is thc (4 x 4) unity matrix, and the quantities
& ?R , S,T stand for the fobwing expressions;

The Hamiltonian matrix of thr rz-valence band can also be derivcd using
the method of iiivarinnts, which was discuss& at an earlier stagc. To do this,
one needs the angular momentum matricrs for spin J = as well aa their 4,
products with earh other. These matrices arp given in Appendix A. The
2.7. k . p +method 197

resulting Bamiltonian matrix agrees exactly with that of equation (2.374).


This means that the neglect of the coupling terms between the spin-orbit and
k p-interactions has no influence on the general form of the r$-Hamiltonian
matrix. Even with this additional approximation one obtains the most gen-
eral I'8+-Hamiltonian compatible with the symmetry of the crystal. Only the
explicit expressions for the constants L , M , N in the matrix DZ:, of (2.343)
are affected. Since these arc understood as empirical parameters, this also
does not play an important role.
The matrix (2.374) has the two 2-fold degenerate eigenvalues E&(k),

(2.376)
2
Using the explicit expressions for Q, T, R , S, this yields

1 1 1
A = -(L
3
+ 2M), B = -3( L - Ad), C 2 7 -[N2 - ( L - M)')].
3
(2.378)

Thc:'T (2 x 2)-matrix block of the full (6 x 6)-matrix is already diagonal.


Its 2-fold degenerate eigenvalue Er7(k) is given by

The energy level Eo1/2/3/4 Bra(0)in equation (2.360) is 4fold degen-


erate at I?. Correspondingly, two 2-fold degenerate bands Erf,(k) arise off
of I?, and starting from the 2-fold ckgenerdte level Ew5/6 = Ey7(0)at r in
(2.362),which lies at an energy separation A below, a similar %fold degen-
cratc band Ep,(k) (see Figure 2.26) evolves off of r. The 2-fold degeneracy
of ixll band6 is a consequence of time reversal symmetry jointly with spatial
inversion symmetry (see Appendix A). The E&-band has weaker curvature,
it corresponds to the band of heavy holas. For k-, = k-, = 0, the pertinent
eigenfunctions are I$$) and . ) ; :1 'l'hc EFa-band is that of the Zzght holes.
-
The eigenfunctions for k, - k, - 0 read) ; ; 1 and ;). ;1 For arbitrary k-
directions, the heavy and light hole states are linear combinations of all four
basis functions ):;I, I:;), . ) ; ;1
3T and
122)
The described structure of the valence band around r represents what
is called the Lallznger-Kohn modal. The particular form of the bands is
determined by the three consttlnts A , B,C. Instead of the latter, one can use
the dimensionless parameters 7 1 , 7 2 , 7 3 called Luttznger parameters, defined
by :
198 Chapter 2. Electronic structure of ideal crystals

I s; II [loo] 5; II [Ill]

I G+l

Figure 2.26: Valence band structure of diamond and zincblende type semicon-
ductors in the Luttinger-Kohn model. The dispersion for the two k-directions is
different (band warping).

h2 ha h2
A -----TI, B = -,2~, C2 = ---12(7,2 - 7:). (23 8 0 )
2m Lni 2m
Values of the Luttinger parameters are listed in Table 2.12. The constants
L , M , N , which were originally used, may be expressed in terms of the Lut-
tinger parameters by the relations:

Both energy band functions E&(k) and Er.,(k) depend on the square of
Ikl, i.e. they are parabolic. This would not have resulted if the interaction
between the rBf-states, and the spin-orbit-split r,f states, had not been
neglected as it in fact was. Concerning the dependencies on the direction
of k, the spin-orbit-split band Er7(k) is isotropic, while the heavy and light
f
hole bands Er8(k) are not. In their case one again has a warping of energy
bands as discussed above (see Figure 2.26). The constant C measures the
strength of warping. In the case C = 0, the warping vanishes. For C # 0,
the point k = 0 is a singularity of the energy band functions Erf,(k) in that
the second derivatives with respect to the components of k depend on the
direction from which one approaches the point k = 0. The effective mass
tensor, as given by equation (2.196), is not defined in such circumstances.
Instead, one can define an anisotropic effective mass, by differentiating in
equation (2.377) not with respect to the components of k,but with respect
to Ikl. If the warping of energy bands is ignored, for instance, by averaging
2.7. k.p -method 199

Table 2.12: k .p-band parameters for selected diamond and zincblende type serni-
conductors. E i , A, and E , in eV. E p = ( 2 m / h 2 ) P 2is a measure of constant P in
Kane's band model. The values for 71, yz, 7 3 are adjusted to the Luttinger-Kohn
model. Temperature below 70 K . (After Landoldt-Bomstein, 1982.)

Material 71 yz 73 Ei A Ep
C 0.94 0.23 0.23 5.48 0
Si 4.26 0.38 1.56 3.4 0.044
Ge 13.35 4.25 5.69 0.90 0.29 26.3
cz - S n 19.2 13.2 16.3 -0.4 0.8 39
GaAs 6.85 2.1 2.9 1.52 0.34 25.7
GaSb 11.8 4.03 5.26 0.70 0.8 22.4
InSb 35.08 15.64 16.91 0.18 0.98 21.2
ZnSe 4.8 0.67 1.53 2.67 0.42
HgTe 12.8 10.6 8.8 -0.30 1.08 17.5

over all directions, this yields the ordinary isotropic heavy and light hole
masses but in the sense of an average.
The Luttinger-Kohn model was described above for the case of diamond
type semiconductors. Formally, for materials having the zincblende struc-
ture, the model does not apply because the matrix elements of the momen-
tum operator p between the triply degenerate valence band states without
spin. IvrnO), are in general non-zero: These states transform according to
the vector representation r15 of the tetrahedral group T d r and the matrix
elements (vrnOlplvm'0) belong to the product representation I'15 x I'15 x r15.
The latter contains the unity representation, as distinguished from diamond
type semiconductors, where the unity representation does not occur in the
corresponding I'i, x r15 x product. The reason for this is the absence of
inversion symmetry in the zincblende structure. The non-vanishing matrix
elements (omOlplvrn'0) give rise to terms linear with respect t o k in the
I'15 valence band Hamiltonian, besides the quadratic terms which are are al-
ready present in the diamond case (see equation (2.344)).However, as a rule,
the k-linear terms are small, and the Luttinger-Kohn model also applies to
zincblende type materials, provided the other requirements which underlie
this model are satisfied. Table 2.12 therefore also contains Luttinger-Kohn
parameters for semiconductors of the zincblende type.
The most important requirement for the Luttinger-Kohn model to be valid is
200 Chapter 2. Electrunic structrrre of ideal crystals

the validity of the assumption that the k p-interaction of the valence band
with thc deepest conduction band is weak enough to be treated by means
of perturbation theory. This is justified as long as the encrgy separation
EF from the lowest condiiclion band and the I&-valenre band (not to be
confused with the fimdamental cnergy gap) is siifficiently large. One expects
deviations from the Luttinger-Kohn model to become noticeable if E i is
sinall Table 2.12 shows thal the EF-value for InSb, for example, clearly
lies below those for C, Si and Ge. In the case of a-Sn and HgTe, Ef; even
becomes negative. Simultaneously, the spin-orbit splitting energy A becomes
relatively large in some of the zincblende type materials, so that even A > /i$
holds. Describing of l h e valence band of such semiconductors by means
o l the Luttingrr-Kohn model would entail treating the effect of the remote
spin-orbit-split band exactly, while taking iuto account only the energetically
closer conduction band by means of perturbation theory. Such a procedure is
not meaningful and one must seek a different, more appropriate description.
A model which is tailored exactly to such circumstances is the Kana
model, which we will now discuss. In this matler, we asbume that the point
group of equivalent directions is the tetrahedral group rh, and no longer the
cubic group o h as above, thercby encompessing both typeb of semiconduc-
lois, those of zinrldende type EM well as those of diamond type. In the latter
case, inversion symmetry still has to be added. This involves a spwialization
of the results, which may be casily done, should thc need arise.

2.7.4 Kane model

The Kane model is based on the following assumptions.


Pzrstly, it is assumed that the k. p-interaction of the valence band with the
deqest conduction band at 1 is so strong that it must bc treated exactly.
Secondly, at r the spiuless valcnce band should have the symmetry T15,and
the spinlesu conduction band should have the symmetry rl. This assump-
tion corresponds to the situation which actually exists in semiconductors of
Llncblendc type.
Thirdly, the interaction of the valence and conduction bands with all re-
maining bands (referrcd to as ~ e m o t is~ )assumed to be srrrall, so that it rriay
be treated by perturbation theory, similar l o the Lultinger-Kohn model in
which the interaction of the valence band with all other bands was treated
in this way. Here, we will simplify further and neglwt this interaction COHI-
plelely. Tn addition, we will also neglect the direct k .p-inleiaction among
the three r15-valence bands which, as mentioned above, does not rigorously
vanish for zincblenrle type crystals, giving risr to k-linmr terms in the Hamil-
tonian. It turns out that the latter approximation is valid in most cases.
2.7. k.p -methad 201

Neglect of interaction with remote bands

We analyzr the generally valid Ychrbdinger equation (2.351) using the as-
sumptions and approximations discussed, above. considering spin md spin-
orbit interaction horn the outset. We again denote the thrw valence band
indices by urn, ni - I.u, z. The conduction band index c will be augmented
by 3 . idirating the 3- or rl-synlm&-y of the conduction band state at r
(the coincidence of this notation s with that of the s p h variable 9 is unfortu-
nate, but unavoidable, and the reader should keep the dxerent meaning of
s dearly in mirid to avoid confusion). The pertinent spin-dependent Bloch
functions at k - 0 are IvmuO) and I c s v O ) , rmpectively. The matrix elements
of the term Ho(k] of f f k X are given by

ri2 2 ,
(vmaO)Ho(k)jvm'dO)= bmm~6umi-k (2.382)
2m
h2 2
(csdlHo(k)lcsdO) = 6,,)-k , (2.383)
2m
(wmOlHo(k)IcsO)= 0. (2.384)
Iii the otkw o p ~ a t o rterm of A V ~ .namely
~, {Fa/m)k-.?i,WP may neglect thc
spin-dqwndcnt part by virtue of the same arguments as in the Luttinger-
kohn model. The three needed matrix elements of this operatw may then
be determined as

(2.385)

The matrix elements (vmOtplzlm'0) of p between the valence band states


are wglectctcd in accordance with the assumptions madr above. The diagonal
rlemenl (csOlplcs0) of p in the conduction band state IcsO) vanishes exactly.
The matrix elements {cs01plvmO) between valence and conduction band
states tranuform according to the produrt representation r1 x I115 x r15
= Fl + trls i l725. Since the unity representation is contained in it
exactly once, the matrix does not haw to vanish; it contains exactly one
independed rlcnwnt. As such, one may chose ( c s O l p , l v d ) and set it equal
t o z ( m / h ) P . The nun-vanishing matrix elements of p are then given by the
relation
202 Chapter 2. Electronic strr1c:ture of ideal crystals

m
~ c s o ~ p , ~=7~~r r8o~~~ p v =
~ I( cv syo~/ p
) , ~ ~=
~ ti .F,-o- )~ (2.386)

The factor 2 giiarantws that P i s real, if the Bloch factors are real as we
assume. The other factor (m/Ta)was introduced for convenience in the final
dispersion rclations. With the moinentiim matrix elements of (2.3861, thp
Hamiltonian matrix (pu01Hk ,(k)lpuO) takes the form

(2.387)

Ec ,iPk, iYk, 0 0 0 0
--iPk, 0 0 0 0 0 0
-iPkg 0 0 0 0 0 0
-iPk, 0 0 0 0 0 0
0 0 0 E C iPk, iPk, iPk,
0 0 0 -iPk, 0 0 0
0 0 0 -iPk, 0 0 0
0 0 0 -iPk, 0 0 I]

where thr lines and columns arp ordprcd in the sequence /s I), la I), Ig I),
I2 th I. 1). lz 1), IY 11, 12 1).
Finally ihe matrix demerits (puO1HsolpmO) nl t h e spiri-orbit interaction
operator ITrn haw to lie detrrminpd. For the Flj-vaknrc band rlernents
(wmcrOlH,,lwmaO), one can adopt the results which were formerly derived
for the valence band. because li5 coincides with I15 for the tetrahedral
group. There are new matrix elPrnents (vmuOlH,lrsaO) and (csuQIH,oj
cscr0) involving the conduction band states. The coordinatedependent fac-
tor of the first matrix element transforms according to the representation
r15 x l?25 x Il - Iz+ I15 +
r25. The unity representation does not oc
cur here, thus this factor vanishes and with it the whole matrix element
(umaOIH,,I csaO), whence

(um.aOlHsIcsrr0)= 0. (2.388)
The caordinatedeyendrnt factor of (csuOlH,olcsdO) belongs to the product
representation rlx r 2 5 x rl :rztj arid must therefore likewise wnish,
2.7. k . p -method 203

Thus, the total Hamiltonian Hk.T + H,, is gken by the following matrix
representation

,
Ec iPk, iPk, iPk, 0 0 0 0
-iPk, 0 -ig
A 0 0 -
0 O ?
-iPkg i$ o o 0 0 0 -27
.A

-iPk, 0 0 0 0 -- $ i+ o
0 0 0 0 Ec iPkz iPk, iPk,
0 0 0 -A
-
3 --iPk, 0 i+ o
0 0 0 -2z .A
-iPkY -+ 0 0
0 a
- .A -iPk, 0 0 0
\
3 2 5 O

Here, the order of rows and columns is the same as in (2.387). For k = 0
and vanishing Ec, this matrix reduces to the spin-orbit interaction opcrator
WSw If one rearranges the rows and columns of this matrix in such a way
that those relating to the conduction band states 1s 1) and 1s J) occur in the
left upper corner, side by side, then the matrix decomposes into a (2 x 2)-
block for the conduction bt~ncl,and B (6 x 6)-block for the valence band.
The eigenfunctioiis of the two blocks arc simultaneously also e i g e h c t i o n s
of the total matrix. T h {2 x 2)-co,nduction band block is already diagonal,
i.c. 1s t) and 1s J] are eigenhnctions of the Hamiltonian matrix (2.390) at
k - 0. The (6 x G)-vdencr band block is identical with the matrix of the
spin-orbit interaction operator H,, of (2.350) for the Luttinger-Kohn model.
The eigenfunctions at k - 0 here are therefore also the vectors 1$ms) and
i$rn 1) of ihe ctngpubr momentum basis (2.366).
T
The latter basis should be particularly suitable for solution of the eigen-
value problem for the Hamiltonian matrix (2.490) at k # 0. The matrix
(2.390) is, however, so simple, indeed, that one can also obtain the secular
equation directly. We will do t h s . before we h r t h w consider the angular
momentum basis. To diminate the free elmtron part ( h 2 / 2 m ) k 2from the
eigenvalues Elk), we write them in the form

h2
E(k) :E(k) -I--k2 (2.391)
2m
where E(k) satisfies the following
204 Chapter 2. Electronic structure of ideal crystals

Secular equation

(2.392)

The fact that the two factors in round brackets appear squared, signifies an at
least 2-fold degeneracy of all eigenvalues. The reason for this is. again, time
reversal symmetry jointly with spatial inversion symmetry (we remind the
reader that the term of the Hamiltonian which can break inversion symmetry
in the case of Td-symmetry has been neglected). Accordingly, one has in
general four 2-fold degenerate bands Ei(k).Ei(k),E$(k),Ei(k). It is also
noteworthy that in the secular equation (2.392). k enters only in the form of
k2. This means that all four bands are isotropic, in contrast to the Luttinger-
Kohn model where a warping of the valence bands occurs. In the case of
k = 0, the energy levels of (2.392) are given by

One may draw conclusions from these expressions in regard to the meaning
of the four energy bands E,(k): El(k) is the J?s-conduction band. E z ( k )
2nd Es(k) are the two upper degenerate rs-valence bands at r. and E4(k)
corresponds to the spin-orbit-split rT-valence band. The energy separation
EL of the r6-conduction band and the I'g-valence bend at. r is obtained as

(2.394)

As long as E: is positive, it represents the energy gap E , at I?. The case of


negative E,' is discussed below. For one of the two upper valence bands - the
one which arises from the vanishing of the first factor of the secular equation
(2.392) and which is denoted by i = 2 - the energy El(k) does not depend
on k. For E$(k).a k-dependence follows with finite negative curvature, as
we will soon see. Thus E;(k) corresponds to a band of (infinitely) heavy
holes, and Ei(k) to a band of light holes. If one adds the (Ti2/2m)k2-term,
then the band Ez(k) displays a positive curvature. It is relatively small
because of the large free electron mass, but the positive sign contradicts
what is to be expected for a valence band. This unexpected prediction for
E z ( k ) results from the fact that the interaction of the valence band with all
remaining bands, except with the deepest conduction band, was completely
neglected. In order to treat the heavy hole band correctly, the interaction
with remote bands must also be considered at least by perturbation theory
as in the Luttinger-Kohn model. This will be done below.
205

Interaction with remote bands

According to the assumptions made at the outset. the interaction with r e


mote bands is weak and may be taken into account by means of second order
k. p perturbation theory. The 8 x 8 Hamiltonian matrix of cquatiori (2,390)
for the conriuclion-valence band complex contains two 4 x 4 blocks of def-
inite spin with rows and columns referring to the conduction band s state
and the three 2--,y--, z-valence band statrs without spin, respectively. To
include the inteTactjori with remote bands. RII additional 4 x 4 matrix o
second order in k has to be added to each of these 4 x 4 blocks. Since the
interaction bctween conduct ion and valrncc b a l d states contributes alrpady
in first order, second order corrections occurring at $2--,sy-, s z - , and
zs--, ys-, zs-positions may be omitted. For the ss-element and the 3 x 3
valence band submatrix, second order corrections beconie import ant. Their
genpral forms follow from symmetry arguments as above. The correction of
the ss-element may be written in the form Ack2, with A , a constant. The
perturbation correction to the 3 x 3 valence band submatrix has the general
form of the 3 x 3 matrix in qiiation (2.344) with parameters L , M. N &fin-1
like the matrix elements D E , D g in equation (2.341), however, with
the conduction band excluded from the summation over { t bewuse this band
is not reinotc. The 4 x 4 matrix block thus determined is added to carti of
the two 4 x 4 diagonal blocks already present in the 8 x 8 matrix (2.390).
Finally, the whole 8 x 8 matrix is subjected to a unitary transformation
into a basis set in which the spin-orbit interaction part of the Hamiltonian
becomes diagonal. The latter requirement is evidently satisfied by a basis
which, as in the Luttinger-Kohn model. contains the 6 angular momentum
eigenfunctions I$:), I$$), I:!),,):;I I%+). I;$), and in addition the two con-
duction band states 1s 1) and 1s 1). This corresponds to a 8 x 8 unitary
transformation matrix composed of a 2 x 2 unity matrix block for the two
conduction band states, and the 6 x 6 matrix block from equation (2.371)
for the six valence band states.
Carrying out the unitary transformation one obtains the general 8 x 8
Kane Hamiltonian which applies to any diamond or zincblende type materi-
als, including those which are already well described by the Luttinger-Kohn
model. However, even in these cases the Kane model is more precise than the
Luttinger-Kohn model, because the valence-conduction band interaction is
treated exactly rather than approximately like in the Luttinger-Kohn model.
If one uses the Kane model in cases in which the Luttinger-Kohn model al-
ready works well. one has to be aware that generally the parameters 71, 7 2 ,
73 have different Values in the two models since those of the Luttinger-Kohn
model contain the valenceconduction band interaction while those of the
Kane model do not so.
The general 8 x 8 Kane Hamiltonian is given by rather lengthy expres-
206 Chapter 2. EIwtronic structure of ideal crystals

sions. To avoid these below, the 4 x 4 block matrix of the remote band
interaction will be reduced to a special case before proceeding further. We
put L = M -- A,, and N = 0, which means physically that the remote
bands affect heavy and light holes in the same way, and do not disturb
the isotropy of the bands. -Ordering rows and columns in the sequence
1s t), 1s l),I);:, I;$), . . . , I$$), the transformed 8 x 8 Hamiltonian matrix
with simplified remote band interaction becomes

U 0 aP,
0 U D -&
-iP- 0 V 0 0 0 0

-fiP- 0 0 0 D 0

-i& -i$&P, 0 v 0 0 O

0 V 0 0

0 0 W 0

0 0 0 U'
(2.3

where the notations l'h = (1/&)P(kz f ik,), Pz = Pk,, U = Ec A&', +


+
V = (1/3)A A,k2, and W = -(2/3)A +
A,k2 are used. The eigenvalues
of this matrix follow from the secular equation

2
i
E'(k) + -A
3
- A,k2
I-i E'(k) +
3

This equation only differs from equation (2.392) in that the factors whose
vanishing define the conduction and valence bands contain, respectively, the
additional terms A,k2 and A,k2.

Solution of the secular equation in limiting cases

The zeros of the h s t factor in (2.396) yield, as seen previously, the I's-band
of heavy holes (i = 2). However, the dispersion relation for it now reads

A ii2
En(k) = -
3
+ A , k 2+ -k2
2m
(2.397)
2.7. k.p -metbod 207

By choosing a negative value for of appropriate magnitude A,, the dispersion


for heavy holes can be brought into agreement with experimental findings.
The zeros of the second factor in (2.396) determine the dispersion of the
Fa-conduction band ( i = l ) ,the rs-band of light holes (i = 3), and the
spin-orbit-split I'T-band (i = 4). For the conduction band, the dispersion
is changed due to the A,k2-term, and for the two valence bands due to
the A,k2-term. But, here, these corrections are added to already existing
strong dispersion terms. We will therefore neglect them in the following, as
we neglected the weak dispersion duc to the free electron term (li2/2m)k2
earlier. Then the eigenvahe equation for the three bands i = I, 3,4reads

- [E'i(k) + $1 P2k2 = 0. (2.398)


This equation will be solved approximately in three limiting cases with re
sped to the order of magnitude relations between the energy gap E i and
spin-orbit splittiiig energy A, as w ~ l las with respect to the s i g n of E,'.
namely firstly for EF >> A, Eg > 0, secondly for EF -CK A, E; > 0, and
thirdly for IEiI < A, E i < 0. The significance of a negative value of E i
will be discussed while treating the third case. All three cases actually oc-
cur in zincblende type semiconductors, as a look at Table 2.12 immediately
shows.
The first case corresponds to materials with wide energy gaps whose va-
lence band complex could be described just as well by means of the Luttinger-
Kohn model; the second case refers to semiconductors with narrow energy
gaps; and the third to materials whose energy gaps vanish.

Case 1: E: z E , >> A, EF >o

We consider energy values Ei(k)in the various bands with energy separations
IEi(k) - &(O)( from the respective band exbrema which are small compared
with EF. For such energies, the conduction band El(k) approximately obeys
the equation

[El(k)- E,]E, - P 2 k 2= 0. (2.399)


and for the two valence bands E3(k) and E4(k) we have

bi(k) - $1 bi(k)
21
+ TA E, + [&(k) + - I: P2k2 = 0, i = 3 , 4 . (2.400)

From (2.399) it follows immediately that


208 Chapter 2. Electronic structure of ideal crystals

(2.401)

and from (2.400) we obtain

2
1 A P2 A P2 8
E3/4(k) = --
2[-
3
+ -k
E, 2]*ij[?f-Kk2]+m' (2.402)

Under the condition (P2/Eg)k2<< A, a parabolic approximation for the


otherwise non-parabolic valence band E3/4(k) is possible, namely

-(2 13)(P2/E,) k2
E3/4(k) = (2.403)
- (2/3)A - (1/3)(P2/E, ) k2.

The structure of the four bands E1/2/3/4(k) is illustrated in Figure 2.27a.

Case 2: Ef = E, << A, Ef >0

The energy values Ei(k) now have energy separations IEi(k) - Ei(O)I from
the respective band extrema which are small compared to A. This means
they can be comparable to E i . For the conduction band and the light hole
band one then gets from (2.398), approximately,

= 0, i = 1,3, (2.404)

from which

follows. In general, the dispersion laws for the electrons and light holes are
again non-parabolic. Only for energy separations from the band extrema
which are small compared to E,, more exactly for P2k2 << (8/3)E,, a k2-
dependence emerges, namely
\i
2.7. k . p -method 209

b)

7
A r,
2

I 4
I 4

Figure 2.27: Valence band structure of zincblende type semiconductorsin the Kane
model for limiting cases: (a) E, >> A, ( b ) E g<< A, E: > 0, ( c ) E , << A, E; < 0.

Since the energy region of width Eg above the band extrema is relatively
narrow for the narrow gap semiconductors considered here, one has in these
materials, even at relatively small separations from the band edges, non-
parabolicities in the dispersion laws of the electrons and light holes. As far
as these particles are concerned, small energy gaps and non-parabolicities
occur together.
For the spin-orbit-split band, one obtains, without further approxima-
t ions,

2 PZ2
E4(k)= --A - -k . (2.407)
3 3A
The dispersion curves of all four bands in the limiting case considered here
are depicted in Figure 2.27b.

Case 3: IEil << A, E i < 0

According to equation (2.394), negative values of EF mean that the r6-


2 10 Chapter 2. Electronic structure of ideal crystals

conduction band level lies below the rglevel. The relation lEil << A
guarantees that the spin-orbit-split r;.-level is found further below it (see
Figure 2 . 2 7 ~ ) .If. again, only energy values are considered with separations
IE,(k)- E,(O)I from the respective band extrema which are small compared
to A, one obtains dispersion relations having the same form as in the previ-
ously considered case EF << A , E , > 0 (see equation (2.406). For electrons
and light holes they yield. under the condition (P2/A)k2 << IEil, the ap-
proximate parabolic dispersion laws

Because of the negative sign of E:, band 1 now lies energetically lower and
exhibits a negative effective mass, and band 3 lies higher and has a positive
effective mass. For the spin-orbit-split band E4(k), equation (2.407) holds
unchanged, and for the band E i ( k ) of the heavy holes, relation (2.397) also
remains the same. Thus, the band E 3 ( k ) is, among all four bands, the highest
energetically with the exception of I', where it degenerates with the band of
the heavy holes. Since the 8 valence electrons of a zincblende type semicon-
ductor are only enough to occupy 6 of the 8 bands of the I?;.--. FS-. ??&band
complex - 2 electrons per unit cell are necessary to fill the deepest I's -valence
band, omitted from consideration here - the E3[k) band remains empty. It
becomes the conduction band, which is also in accord with its positive effec-
tive mass. At r, its separation from the uppermost valence band, the Ez(k)
band of heavy holes, is zero. This means that the energy gap vanishes in this
case. Negative d u e s of the parameter E i . which signifies the energy gap
when it is positive, cause EF to lose this significance, and the real energy
gap becomes zero. Materials with vanishing energy gap are called zero-gap
semtconductors. Examples are HgTe as a zincblende type semiconductor,
and D - SR as a semiconductor of the diamond type.
In Table 2.13. the effective masses m t are listed for the three limiting
cases considered above. Generally, one has 1m.I I 5 I rn31 < IrnZl. The ef-
fective masses of the electrons and light holes are proportional t o / E g l / P 2
throughout. The rule discussed above for degenerate bands is thereby con-
firmed to be valid also for non-degenerate bands wherein mass decreases as
lEgl decreases and P increases. For IE,l << A the effective masses of the
electrons and light holes are almost identical, for E, >> A the electrons are
lighter than the light holes. For the spin-orbit-split band, the proportion-
ality of the effective masses to IE,l/P2 exists only in the case E, >> A; for
<< A the mass mS becomes independent of Es and proportional to A.
2.8. Rand structure of important seniiconductors 211

Band EF>>A,E;>O EL<A,h',r>O EL << A, EF < 0


rdl) 1.(F,/P2) (3/2). (E,/P2) -(3/2) ' (I 8; I / P 2 )

rs(3) -(3/2) . ( E g / P ' ) -(3/2) . ( E S / f " ) (3/2) . (I -Ei I / p 2 )

rd4) -3 ' (E,/P2) -3 ' (A/Fs,) -3 ' (A/ I q I)

2.8 Band structure of important semiconductors


In this section we discuss the band structures of some important semicon-
ductors. In all cases, the results presented are based on both theoretical and
experiment a1 investigations.
Experimental data concerning band struct ure of semiconductors arc mainly
obtained by means of optical reflectance spectroscopy. It turns out that char-
acteristic structures of the reflectance spectra, like peaks or shoulders, are
directly related to optical transitions at critical points of the energy difference
between the initial and final bands involved. The frequencies of these struc-
tures are experimental measures of the energy separations between initial and
final bands at critical points, To enhance the charactelistic spectral features,
changes of the reflectance spectra are measured due to external perturba-
tions, as, for example mechanical strain, clmtiic and magnetic fields, light,
or heat. By modulating the perturbations periodically in time with frequen-
cies in the kHz range, these spectral changes can be measured very precisely
by means of frequency and phase scnsitive techniques. Exttniple5 of this so-
called modulataon spectroscopy are electrorepectance (ER), pzezorejlectanw,
thermarejlectance, and photoreflectance (PR) (for morc on elect roreflwtance
see section 3.7). Details of the band structure at critical points, like effective
masses of free carriers, may also be extracted from transport meaxmemenits.
Again, external perturbations, in particular, magnetic fields, are applied to
induce changes. Magnetotransport phenomena, like rnngnetoreuzutanc.r and
Shubnzkov-dp-Haas effect are examples. In cyclotron rmonance one mea-
sures the absorption of microwave radiation by a semiconductor sample in
the presence of a magnetic field to obtain the effective masses of electrons
and holes (section 3.8).
None of the experimental methods is capable of revealing the entire band
strurture of a given semiconductor material at all poiutv of the first BZ. To
212 Chapter 2. Electronic structure of ideal crystds

obtain this. one is nhliged to carry out band structure calculations. Expai-
meutal data enter thrse calculatiolis in various ways. Tlus is obvious if em-
pirical methods are cmployed their results have to be fitted to experimental
data, as, for example, to energy separations between bands at critical points
obtained from modulation spectroscopy. Less obvious. but nevertheless ex-
isting, is the need of experimental data for ab-initio calculations. Although
these methods are free of fitting parameters, various approximations are in-
volved which call for experimental confirmation or even corrections of the
results. as, for example, in the case of the erroneous fundamental energy gap
in the local density approximation.
Below we represent the results of of uumprical hand structure calcula-
tions using one or mother of the methods described in section 2.5. We will
not specify which particular method was applied since that is not of interest
here. Our main concern is with the qualitative features of the energy bands.
We will demonstratp that thew may be understood. at least partially, just
by means of the general results derived in the preceding sections and in Ap-
pendix A. This is particularly true for features irivolving the degree of band
degrnrracy at symmetry points of the first B Z , which follow from the i r r e
dudble representations of the space group of the crystal under consideration
(see Appendix A). The band structure models derived by the empty lattice
approach. the k . p method, and the tight binding method in sections 2.4,
2.6 and 2.7, respectively. will also be helpful. We begin our discussion with
silicon.

2.8J Silicon
In Figure 2.28 the band structure of Si is shown. Spin-orbit interaction plays
only a minor role for the ovcwJl behavior of energy bands in Si, 80 the energy
levels may be classified by means of the ordinary irreducible representations
of the small point groups of the wavevectors k. l'hesr are subgroups of the
full point group 01, of eqiiivalent dirrctions. According t o what we already
know about the dimensions of these representations at the various symmetry
points of the first B Z (see Table 2.5 and Appendix A), one expcrts at most
%fold degenerate levels at the symmetry center I', only 2-fold a t the the
symmetry point X. and at most 2-fold on thp symmetry lines A, A and at
L. This expectation is codinned by the band structure shown in Figure
2.28. The deepest energy level of the entire band structure occurs at r, i s
non-degenerate. and belongs to the irreducible representation r'1. The same
result was previously obtained in the band structure analysis of diamond
type crystals by means of t h e empty lattice altd tight binding approaches in
earlipr swtions (see, Figwrs 2.13 and 2.2L rcspectively). Also, off of l?, the
numerical calculation of the PI-band of Figure 2.28 exhibits the behavior
prdictfd by these simple approaches. DiEmmrw between the numerical
2.8. Band structure of important semiconductors 2 13

Figure 2.28: Rmid structure of Si. The energy unit i eV. (After Chelikowskp and
Cohen, 1974.j

and empty-lattice band structures occur at the second level at as may ber
seen from Figure 2.29. The second empty lattice has an 8-fold degeneracy
which exceeds what is compatible with the Oh-symmetry of r.
In a real
diamond type crystal this degeneracy is removed. As indicated in Figure
2.29, a splitting into levels of rl-,I'b-, I'b5-, and r l 5 - s y m e t r y will occur.
In this way the corresponding rk5-, rls-, and I'k-levels in Figure 2.28 could
have emerged from the 8-fold degenerate empty lattice level. The tight
binding analysis of the band structure of diamond type semiconductors in
section 2.5 has already shown that the second level (from below) at I? has
symmetry For the third level at r, the representations r15 or were
possible, according to this analysis. 4 s Figure 2.28 shows, r15 applies in the
case of Si, while the I'b-level is the fourth.
A410ngthe A- and A-lines, the two 3-fold degenerate let~lsI'L and r15
must split since only 1- and 2-dimensional irreducible representations are
possible there. Both a %fold splitting into 3 simple bands as well its a 2-fold
splitting in a 2-fold degenerate and a simple band are conceivable. From
Figure 2.28 one sees that the latter case holds, both along the A and A-
lines. and for the I?&- and the I'ls-levels. Since there are only 2-dimensional
representations at X (for an explanation see Appendix A), two simple bands
along the A-line must merge at this point, as actually happens in Figure 2.28.
214 Chapter 2. Electronic structure of ideal crystals

Figure 2.29: Band structure of


the empty fcc lattice. The energy
unit is the same aa in Figure 2.28.
t lo
A
0
The irreducible representations of -w 8
the energy bands are also indicated. m
L
For comparison, the band structure cu
C
of Si is shown in the same figure by w 6
dotted lines.

r, 62 5;
r; 5' 5 5 6;

r1
r A
Wavevector - X

A 2-fold degenerate band along A cannot merge with another band at X , but
miist terminate in a doubly degenerate level at X. From this it follows, for
example, that the upper of the two bands arising from the I'k5-level along the
A-line must be 2-fold degenerate, and the lower band must be simple. A look
at Table 2.4 shows that the only 2-dimensional representation of the small
point group of A is A5. At X this representation is compatible either with
X 1 or X 3 (the compatibility relations between irreducible representations
are derived I n Appendix A). Figure 2.28 shows that X 1 is correct in this
case. In a similar way one may conclude that the representation of the lower
level at X ,arising from the I'i,-level, must be X z . A similar analysis for the
splitting of the rls-level along the A-line shows that the lower level is non-
degenerate and belongs to the irreducible representation A l , and the upper
level is 2-fold degenerate and belongs to .As. The intersection between the
Ah-band emerging from the r!+vel, with the As-band emerging from the
l'ls-level, is not due to symmetry, but reflects an accidental degeneracy.

Consider next the two bands arising from the l'$s-level on the A-line.
The upper is the 2-fold degenerate As-baud, and the lower is the simple
A1-band. The simple band does not merge with another simple band at
L , but remains separated from it by a finite gap, in contrast to the band
behavior at X. This is possible because at L there arc also 1-dimensional
representations, L1 and L i , which give rise to the two lowest levels at L .
The splitting of the rls-level along the A-line is quite similar to that of the
l?~5-level;the lower band is non-degenerate and belongs to A l , and the upper
2.8. Band structure of important semiconductors 2 15

is 2-fold degenerate and belongs t o h ~ .


The occupation of the energy bands of Si and other diamond type crystals
has already been discussed in section 2.6, using the band structure which
follows in tight binding approximation. In the ground state of the crystal,
the band associated with the 1-fold I'&vel, and the bands arising from the
%fold l?h5-1evd, are completely occupkd, while the above-lying r16- and rb-
bands are completely empty. Thus the rl- and r&-bands form the valence
bands of Si. and the rls- and rb-bands. as well as all higher bands form the
conduction hands. The two groups of bands are separated by k-dependent
forbidden energy regions, marked by dashes in Figure 2.28. Moreover, there
are forbidden energy values which occur at all k-vectors. This means that
the bami structme calcidations for Si yield a finite energy gap, and, inded,
they explain the semiconductor character of Si.
While the absolute maximum of the uppermost valence band lies at r,
the absolute minimum of the lowest conduction band is located on the A-line
close to the edge of the h s t B Z , and its irreducible representation is AL-
Semiconductors with the absolute extrema of the valence and conduction
bands located at different points of the first B Z , are called ipadiwct. Silicon
is, therefore, an indirect aernicondurtor. If the extrema occur at the same
point, the- semiconductor i s called dawct. The property of a semiconductor
material in being direct or indirect (Table 2.14 contains information about
this) has important physical and technological consequences. For example,
indirect materials are in general not suitable for manufacturing light-emitting
devices, unless one takes special measures such as a particular doping.
Following the general lines of section 2.5, we will now examine the efec-
tive mass tensors of Si. We will restrict oursehes to the two critical points
mentioned above, in which the conduction band has its absolute minimum
and the valence band has its absolute maximum - the vicinities of these
points are the regions of the first 32 which. under thermodynamic equi-
librium conditions, host most of the free eIectrons and holes. The effective
masses at these points are therefore the effective masses of the electrons and
holes of silicon.
Owing to the cubic symmetry of the band structure of Si, a minimum of
the conduction band on a particular A-line is automatically accompanied by
5 other minima on the star of A-lines. This means that there are, altogether,
6 minima OT valleys, a term which is often used for the vicinities of the min-
ima. Since Si has several valleys. one calls it a many-valley semiconductor.
The valleys are centered at the points

with kc = 0.85(2n/a). The cubic axes are, simultaneously, the principal


2 16 Chapter 2. Electronic striictnre of ideal crystals

axe8 of the effective mass tensors of the various valleys. Each principal axis
represents a Cfold rotation symmetry axis. To p r o r d further. we select,
arbitrarily, the valley centered at (O,O, kJ. The band structure of this valley
is given by the expressions (2.201) and (2.2@2), which are applicable here,
as their conditions of validity are satisfied. Setting the band index v equal
to c, which refers to thp coiiductioii band, the dispersion rdation EJk) of
this band becomes

(2.409)

If t,hc zero point of the energy scale is put at. t h r valence band maximum, as
we do here. then E , is t,he fundamental energy gap.
In the vicinity of the valencc band maximum, the general results of the
k . p-method of section 2.7 are applicable. Without spin-orbit interaction
(see cqiiations (2.345) to (2.348)) one has two valence bands for each of
the two symmetry directions A and A, an upper E,l,z(k) which is 2-fold
degenerate, and a lower &(k) which i s non-degenerate. Along less sym-
mdricnl k-directions E,l/z(k) split 5 into two bands. However, spin-orbit
interaction cannot be neglected; although it has little effect on the overall
band structure of Si, it inftuences the valence band structure considerably
in an energy interval of several kI' below the rnaxiniurn at r, which is the
energy region where most of the holes B r e located. Spin-orbit interaction
makes the uppermost valence band level at r, which has a 6;-fold degener-
acy in Si if spin i s considered, split into an uppm &fold degenerat.e I?$-level
and a lower 2-fold degenerate I';-lcvel. The splitting rncrgy amounts to
44 m e V . Away from I?: the upper ra-level decomposes into the two bands
Efp(k) of heavy and light holes according t o equation (2.377). The lower
r$-levd givw rise to the spin-orbit-splitr;-band. The heavy and light hole
bands are strongly warped and each exhibits %oldspin-degeneracy, which
is due to timc reversal syrrirriet,ryin coinbination wihh spatial inversion (see
Appendix A). If warping i s neglected the two bands Era(k)can he described
by isotropic effective mass# m$. 'I'hey are negative because of the maximum
at .'I Extracting the negative sign, we define the positive effective masses
rnEh and m.G1of, respectively, heavy and light holeb;: setting m:,,= - m ; , and
ntl = - m l . For the heavy and light hole valence bands EF8(k) G E,h(k)
and E&(k) = E,,,(k): we h a w

(2.410)

Numerical values for the effective hole masses of Si are given in Table 2.14.
A vivid view of the conduction and valence band structure in the vicinity
of the band edges is provided by the corresponding iso-enerm surfaces. These
2.8. B i d structure dimportant semiconductors 217

Figure 2.30: Iso-energy surfaces of the


electrons (on the left) and holes (on the
right) for Si.
I

are obtained by k i n g a specific energy and then drawing all k-vectors for
which the bands of equations (2.409) and (2.411) yield this energy value (see
Figure 2.30). For the conduction band, the iso-energy surfaces are ellipsoids
of revolution, pointing in the direction of the symmetry axis. Each of the six
star points is the center of such an ellipsoid. For the valence bands, within
the isotropic approximation, the iso-energy surfaces are concentric spheres
centered at I?. The inner sphere corresponds to the light hole band, and the
outer to the heavy hole band. In reality the valence bands are not isotropic
but have only cubic symmetry. Thus their iso-energy surfaces are warped, as
shown in Figure 2.30. If the conduction band is populated by electrons up to
a given energy, then the k-vectors of these electrons lie within the ellipsoid
of revolution corresponding to this energy. Accordingly, the k-values of the
holes lie within one of the two spheres or the two bodies bounded by the
warped surfaces.
218 Chapter 2. Electronic structure of ideal crystals

Figure 2.31: Band structure of Ge. The energy unit is eV. (After Chelikowsky
and Cohen, 1.974.)

2.8.2 Germanium
In Figure 2.31 the band structure of germanium is depicted. It is similar to
that of silicon, owing to the fact that both materials have diamond struc-
ture. The differences between the two band structures are, apart from other
reasom, due to the fact that the spin-orbit interaction is considerably larger
in the case of Ge as compared to Si (see Table 2.14). This results from the
larger orbital velocity of the valence electrons of Ge because of the larger
atomic nucleus of this element. With the regard of spin-orbit interaction
the valencp band maximum at r is formed by an upper 4-fold degenerate
rsf-level and a lowcr 2-fold degenerate I';-level, separated from the upper
level by 340 meV. Away from I',the upper I',f-level decompoaes into the two
2-fold degenerate bands of heavy and light holes of, respectively, A6 and A7
symmetry, As in the case of Si, the bands are warped but often considered
in an isotropic approximation.
The lowest conduction band level r, at 2! arises from the ri-level without
spin, which was the second conduction band level in the case of Si. In Ge this
level i s shifted down considerably so that it becomes lower than the r, and
,?I doublet which arises from the I'15-level without spin. The most important
2.8. Band structure of important semiconductors 219

Table 2.14: Characteristic data of the barid structure of selected semiconductors.


Energies in e V , effective masses in free electron masses. Temperature 300 K. For
Si, Ge, GaAs, Gap, CdTe, and IIgTe read i = u h , v l (heavy, light holes), and for
CdS (hcxagonal phasc), PbTe, Te, and Se read i ~ 1 1 I , (parallel, perpendicular to
syniniebry axis). ( A f t e r Candoldt-Blirnstein, 1982.)
-
-
Milate- C-Blmd- Energy

i_
rial
.
Mimirnurn Gap hl
I_
II
Si A 1.1 1 0.044 0.54 0.15
GI2 L 0.66 0.29 0.3 0.044
GaAS r 1.43 0.34 0.45 0.087
GaP X 2.27 0.08 0.67 0.17
CdTe r 1.43 0.7 0.4 0.1

HPn: r 0.00 0.30 0.3 0.03


CdY r 2.50 0.064 5 0.7
PbTe L 0.30 0.24 0.028 0.31 0.022
k H 0.33 0.07 0.1 0.11 0.24
Z(?) 1.8
-
Se
- -
- -
- -
-

difference between the conduction bands of Ge and Si is the location of the


respective absolute minima. In Ce it does not occur on a A-line, but at L ,
being the B Z boundary point of the A-line. Therefore, the conduction band
of Ge has 8 half-valleys instead of 6 full valleys in the case of Si. The minima
at L are L;-levels arising from the &-band on the A-line which enters the
r7f-level at I.

2.8.3 111-V Semiconductors


For the two 111-V compound semiconductors GaAs and Gap, the band struc-
tures are shown in Figure 2.32. Their first B Z s are the same as that of Si
and Ge, since they have the same Bravais lattice. The full point group of
equivalent directions is T d for both materials. Spin-orbit interaction is im-
portant for the overall band structure of GaAs, while for GaP it may be
neglected. Thus the band structure of GaAs is described in terms of spinor
representations, and that of GaP in terms of ordinary representations. For
both materials, the valence band maximum lies at r.
Without spin and
220 Chapter 2. Electronic structure of ideal crystals

Figure 2.32:Band structure of 111-V semiconductors GaAs (top) and GaP (hot
tom). The energy unit is el. (AfLer Chelikousky rand Gohen, 1974.j
2.8. Band structure of important serniconductors 22 1

I 1 I
i r x U,K r -61
L A rI A XU,K E rI

Figure 2.33: Band structure of IT-VI sernimnductom Cdk (left) and IIgTe (right).
The energy unit is el.. (After Chadi, Walter, Cohen, Patroff and Balkanski, 2972.)

spin-orbit interaction (the case of Gap), it belongs to the irreducible repre


sentation Il5 of the point group Ta which arises from the representation ,;?I
of Oh if the latter is taken as a representation of Td. With spin-orbit interac-
tion (the caw of GtrAs), the r15 vulence band maximum splits into an upper
r8-level, and a lower I7-levd In the upper lg-level, the two 2-fold degener-
ate heavy and light hole bands mcrgr together as in the case of Ge and Si.
The lower l;.-level gives rise to the 8-fold degenerale spin-orbit-split band.
The %fold degenpracy of the valence hands at X,observed in the case of Si
and Ge, splits in CaAs and Gap. The reason or this is that the t,wn atoms
of the primitive unit cell are no longer identical, which means that the point
symmdry of equivalent directions is rduccd from Of, t o Td. Therefore, one
also has 2-dimensional spinor representations ( X s , X:) instead of only one
4-dimensional ( X , ) in the case of Si and Ge,and also 1-dimensional ordinary
representations (XI X2,X3, Xq) instead of only 2-dimensional ones ( X i ,
X z ! x3>Xq) in the case of Si and Ge (see Appendix A).
The conduction bend minirnnrn of GaAs occurs at the r-point and be-
longs to the spinor representation r7. Thus, CaAs has a dirwt energy gap.
One of the peculiarities of its condiiction band structure is the relative min-
imum at the L-point only about 0.4 eV above the absolute minimum atr r.
In the case of Gap, the conduction band minimum occurs at X and belongs
to the representation XI. Since the valence band maximum resides at I,
GaP is an indirect semiconductor.

2.8.4 11-VI semiconductors


The band structures of two typical IT-VI semiconductors with zincblcnde
structure, CdTe and HgTe,are shown in Figure 2.33. The band structure of
222 Chapter 2. Electronic structure of i d 4 crystah

-2

CdTe is similar to that of GaAs, in particular the conduction and valence


band edges are located at the r-point, as also occurs in GaAs. The same
holds for HgTe, but as mentioned in section 2.7, the two bands touch each
other at r, so that the energy gap vanishes. The reason for this is easy to
understand, and was also already discussed in section 2.7: The r6-level which
lies above the Is-level in CdTe, is found below it in HgTe. The 8 valence
electrons per primitive unit cell therefore suffice t o occupy only one of the two
rs-bands merging at I?. The second band remains unoccupied and becomes
2.8. Band structure of important semiconductors 223

I
L r X
lxc,

I
4
K.U r

Figure 2.35: Band structure of PbTe. The energy unit is eV. (After Martinez,
Schliiter and Cohen, 1975.)

the conduction band. Since it is degenerate with the uppermost valence band
at I', the fundamental energy gap is zero. According to the general definitions
of Chapter 1, this means that HgTe is no longer a semiconductor, because
in semiconductors the valence and conduction bands must be separated by
a finite energy gap. It is more fitting to term HgTe a semimetal.

As an example of a 11-VI semiconductor which does not exhibit the


zincblende but rather the wurtzite structure, we chose the hexagonal CdS
(see Figure 2.34). The first B Z is that of the hexagonal Bravais lattice,
which is shown along with its symmetry points in Figure 2.12. CdS has a
direct energy gap at r. Furthermore, the valence band is additionally split
off at r as compared to the cubic materials of diamond and zincblende struc-
ture, because of the hexagonal deformation of CdS. Thus, the valence band
is composed of 3 bands, namely r7-band which is separated from the other
two by spin-orbit interaction, and the two bands r6 and I?7 into which the
rg-band decomposes under the hexagonal deformation. The latter effect is
called crystal field splitting (see the right part of Figure 2.34).
224 Chapter 2. Electronic structure of idea[ crystals

12

-4

-8
r z H K r 1 - z H K r
Figure 2.36: Band structure of Te (left) and Se (right). The energy unit is eV.
(After Maurhke, 1 g71; Stuff, Maachke and Laude, i973.)

2.8.5 IV-VI semiconductors


The band structure of a typical IV-VI semiconductor, PbTe, is depicted in
Figure 2.35. It crystallizes into the rocksalt structure. Its Bravais lattice
is, thus, again the fcc lattice and its first B Z is the same as that of the
diamond and zincblende type crystals. The conduction band minimum and
the valence band maximum both lie at the edge-point L of the first B Z . This
means that PbTe is a direct gap semiconductor with many valleys, a special
feature which is hardly realized in any other semiconductor families apart
from the IV-W compounds. The energy gap of PbTe is relatively small, it
amounts to about 0.2 e l F . Like InSb of the 111-V-compounds, PbTe belongs
to the group of narrow gap semiconductors.

2.8.6 Tellurium and Selenium


The band structures of tellurium and selenium are shown in Figure 2.36.
Both materials have hexagonal Bravais lattices. Therefore, they have the
same first B Z s as the hexagonal CdS. In the case of Te. both the valence
band maximum and the conduction band minimum occur at the H-point of
the first B Z . The material is therefore direct. Se has an indirect gap, the
two edges both lie outside of I-, that of the conduction band is probably at
2 , and that of the valence band is probably at H.
22 5

Chapter 3

Electronic structure of
semiconductor crystals with
perturbations

Real semiconductor crystals are imperfect. They differ from ideal crystals
in their atomic structure as they contain structural defects, and their chemi-
cal cornpodmn is not exactly lhal of the correspondiug ideal crystals. It is,
however, not the atomic structure which provides real semiconductor crystals
with special scientific and technological importance. but their peculiar elec-
tronic structure that is manifested in pronounced macroscopic effmts such
as. for example, the increase of electrical conductirity by orders of magni-
tude. With some justification one can even say that the particular scientific
and technological importance of semiconductors rests on the peculiarities of
the electronic structure of amperfeet semiconductor crystals. Nevertheless,
one must also deal with the atomic structure of imperfect semiconductor
crystals since that is, ultimately, the source of the particular nature of their
electronic structure. This will be done in section 3.1. The one-electron
Schrodinger equation of imperfect semiconductor crystals will be derived in
section 3.2. - 4 important
~ theoretical method or its solution, the so-called
effective mass theory, will be developed in section 3.3. This equation allows
one to determine the effects of perturbations whose interaction potentials
with electrons are slowly varying over the atomic length scale. Interaction
potentials of this kind apply to many perturbations of atomic structure, in
particular to the so-called shallow centers which will be treated in section
3.4. There are. however. also other perturbations. like the so-called deep
centers, to be dealt with in section 3.5, in which potentials exhibit consider-
able change over the atomic length scale. They cannot be treated by means
of effective mass theory, but require other methods. This is also true for
clean semiconductor surfaces considered in section 3.6.
226 Chapter 3. Electronic s-tnictirre of semiconductor crystals with perturbations

Besidc the study of semironductor crystals having a perturbed atomic struc-


tiire, 11 is also of interest to study semiconductors which are exposed to
macroscopic external perturbations such as dectrir OT magnetic fields, me-
chanical strain, or arbifmid atomic superstructures. The electronic structure
of semiconductors in the presence of such macroscopic external perturbations
is treated in the second part of this chapter. This will include superlattices
(section 3.7) and semiconductors subject to an electric field {section 3.8) or
a magnetic fidd (section 3.9). In all cases, again, the effective mass theory
plays a major part in our theoretical understanding. This theory applies
because the macroscopic perturbation potentials are naturally smooth over
the atomic length scale.

3.1 Atomic structure of real semiconductor crys-


t als
The reference system for the description of the atomic structure of a real
crystal is the ideal crystal. As discussed in Chapter I, the latter may be
characterized as follows: All regular sites

Ri 7 rial + r2a2 + 7'3a3 + 6 (3.1)


defined by the crystal structure, are occupied by atoms of the 'correct' chem-
ical species, and other sites are empty. In a real crystal, this characterization
holds for the vast majority of regular and irregular sites - providing the jus-
tification to speak of a crystal at all, albeit a perturbed one. However, there
is deviation from the ideal occupation at some of the regular and irregular
sites. In this sense. one then refers to the crystal as either perturbed or real.

3.1.1 Classification of p e r t u r b a t i o n s
The perturbations to be treated can be classified in accordance with several
points of view. On the one hand, one can distinguish them on the basis
of whether they are of a purely chemical nature, or purely structural, or
mixed type. In the first case, only regular sites of the crystal are occupied,
but. not with chemically 'correct' atoms throughout. One refers t o this as
chemical or cornpositzonal dzsorder. In the case of elemental crystals, this
kind of disorder necessarily means the presence of impuaty atoms. In this
context the perturbed crystal, which hosts the impurity atom. is called the
host crystal. A crystal formed from a chemical compound may also exhibit
compositional disorder without impurity atoms, namely because of a per-
turbed stoichiornetrical composition. In the case of structural perturbations,
only chemically 'correct' atoms are present, but these are not always posi-
tioned on regular crystal sites but also on irregular ones. Furthermore, not
3.1. Atomic structure of red semicoriductor crystals 227

all regular siles are occupied by atoms, some sites remain empty. Structural
perturbations are also called strurtural & f w h or simply dpferts. The third
case is the most general one - regular or irregular sites of the crystal are
partially occupied by rhrrnirnlly wrong atoms, end also the correct atoms
partially occupy wrong sites.
Deviations from an ideal crystal may, on the other hand, be distinguished
according to the macroscopic extension of the perturbations. A perturbation
that is limited to one or a few neighboring regular or irregular crystal sites
is called 0-dimensional or a poznt perturbatzon. If the perturbation extends
over sites located on a line or a planc, it is referred to as a 1-dimensional
or lzne perturbatton and a 2-dimensional or planP perturbatton, rpspert ively,
Combining the two classification schemes, one may refer to structural point
or line perturbations, compositional point perturbations etc. The dimen-
sions in the second classification scheme apply to the microscopic core of
the perturbation. There may be smaller perturbations induced by that core
which extend in three dimensions. Examples are charged impurity atoms,
which, due to their long-range Coulomb forces, change the potential energy
of an electron even over distances large compared to the lattice constant.
Below we will characterize the various perturbations in more detail, start-
ing with point perturbations.

3.1.2 Point perturbations

Types of point perturbations

In this subsection, we describe the most important compositional and struc-


tural point perturbations of semiconductor crystals. An illustration is given
in Figure 3.1.
(1) We begin with an impurity atom on a regular crystal site. Since the
impurity atom substitutes an atom of the host crystal (see Figure 3.lb)
it is referred to as a substztutzonal vrnpuraty. Examples of substitutional
impurities are a phosphorus atom in Si on a Si-site, or a sulphur atom in
GaAs on a As-site. To avoid a somewhat cumbersornc description, in the first
case one uses the symhol S I : P , and in the second, the ~ynibolGaAs ; S A ~ .
As a rule, impurity atoms which are chemically similar to atoms of the host
crystal, are incorporated substitutionally. For this reason, many elements
of the mein group of the periodic table, if added l o group-IV elemental
semiconductors as well as binary 111-V and 11-VI-compound semiconductors,
form substitutional impurities. The substitutional incorporation, in most
cases, occurs on that lattice site which corresponds to the chemically most
similar of the two atonis in the binary compound semiconductor. lherefore,
the doping of GaAs with S leads to the above mentioned point perturbation
228 Chapter 3. Elcctronic structure of semiconductor crystah with perturbations

C a A s : S A (and
~ not GaAs : S G ~ )and
, the doping of GaAs with Si leads to
G a A s : Sica (and not GaAs : S ~ A ~ ) .
(2) An unoccupied regular crystal site is called a vncuncy, as depicted in
Figure 3. Ic. Tn semiconductors made of binary chemical compoiinds, one
has to distinguish between cation and anion vacancies, as shown Figure
3 . 1 ~ .Vacancies occur in all important semiconductor crystals; the general
symbol is V. The vacancy in Si is denoted by S i : c': the cation vacancy in
C h 4 s by G a A s : T;&, and the anion vacancy by GaAs : VA*.
(3)If the impurity atom does not occupy a repular crystal site but a site
between regdar ones, one has an interstitial ampurity atom (Figure 3.16).
In order for an impurity atom to stay at an interstfitrialsite, it must have
sufficiently low energy there. It is quite clear that this will be satisfied for
interstitial sites which either haw high local symmetry or which lie on a
bond between two at'orns. In thc latter case t,he crystal has bond centered
interstitiah. The high symmetry interstitial sites in t e t r a h d a l seniiconduc-
tors may be such with tetrahcdral local symmetry in the neighborhood of
the cation or of the anion (for group IV elemental semiconductors the latter
distinction is void, of course). Moreover, there are high symmetry sites with
hcxagonal symmetry. One refers to these7 respectively, as f e t v u h c h l and
hexagonal interatitiak The incorporation of impurity atoms on interstitial
crystal sites is especially likely when the impurity atom deviates relatively
strongly from the atoms of the host crystal as! for example, in t,he case of
transition metal atoms in semiconductors composrd of elements of the main
groups. The general symbol for an interstitial is I . An interstitid Fe atom
in Si on a tetrahedral site is denoted by Si : I;,.
(4) If a chemically 'correct' atom of the crystal occupics an interstitial site
rather than a regular one, one has a self-interstitial (as shown in Figure
3.le). In order for such a structural point defect to develop in a crystal,
there must h e enough space between the host ntmns, i.e., the crystal should
not be packed too densely. This happens, for example, in the case of tatra-
hedral semiconductors, particularly in Si and Ge which have purely covalent
b ondinE.
( 5 ) If,a crystal consists of two different chemical elements, then an atom of
the h s t may occupy a regular site of t h e sccond, and vice vcrsa Such point
perturbations are called antisite defects, as i l h s t r a t d in Figure 3 . 1 In
~ the
case of CiaAs, for example, a Ga atom may be located ttt an h - s i t e this is
~

called B I ~As-antisite defect,, and tla As atom may occupy a Cia-site - this is
called a Ga-antisite defect. The symbols are As& for the As-antisit>edefect,
and G U Afor ~ the Ga-antisite defect.
Interstitials, vacancies and antisite defects are structural point perlur-
bations, or point defects. The compositional point perturbations, i.e. the
3 . 1 . Atomic structure of real semiconductor crystals 229

Ideal crystal

A -Atom
4
0 B-Atom
Impurity atom

Substitutional impurity (S) SA SB

Vacancy (V)

Interstitial ( I 1

Interst. impurlty - ___


A,-Antisite B,-Antisite

e)

Figure 3.1: Illustration of the most important point perturbations in semiconduc-


tors using the example of a crystal with two atoms per unit cell of the same chemi-
cal element (left-hand side) and different chemical elements (middle and right-hand
side).
230 Chapter 3. Electronic structure of semiconductor cxystals with perturbations

Table 3.1: Electron corifiguratiori of main group elements. In the rightrnost column
the respective closed shells are indicated.

impurity atoms, need to be further specified. This will be done next.

Classification of impurity atoms

The division of elements into groups, which is commonly used in chemistry,


also proves to be helpful for the classification of impurity atoms in semi-
conductors. This is not surprising because the incorporation of an impurity
atom in a crystal indicates a more or less strong chemical bonding. We
summarize this group division of chemical elements below.
The periodic table consists of two types of groups of elements, the main
groups, and the transition groups. Of the first 98 elements, 50 belong to
the main groups and 48 to the transition groups. The elements of the main
groups in Table 3.1 have in common the feature that electron shells with
angular momentum quantum numbers 1 2 2 either do not occur at all or, if
they exist, they are completely filled or completely empty, i.e. no partially
filled shells of this kind occur. The energetically highest, and thus in general,
only partially filled shells of these elements either have 1 = 0 or 1 = 1.
Therefore, they are s- and p-shells. Because of the relatively large spatial
extension of s- and p-shells in comparison with d- and f-shells, the former
are simultaneously also the outer shells of the atoms which are responsible
for chemical bonding. One thus also speaks of sp-bonding elements. The
rare gas elements are special cases, in which the s- and pshells are also
completely occupied.
For the elements of the transition groups presented in Tables 3.2 and
3.3, the shells with 1 2 2 are energetically the highest and, thus, in gen-
3.1. Atomic structure of real semiconductor cry&& 231.

Table 3.2: Electron configuration o transition dernent,s. In the rightmost column


the respective closed shells are shown.

n Iron m o w
ziSc =Ti 23V z4Cr wMn mF'e z7cO Ni
3d4~3~ 3d248' 3d34s2 3d5& 3d"4s2 3d'4s2 3d74s2 3d64aa 3a23p6

era1 they are the not completely occupied oms. Because of the relation
R 2 I +
1 between the main quantum number ?a and the angular mornen-
turn quantum numbe1 E. the d-shells (I - 2) are possible only for n >_ 3.
the f-shells ( 1 - 3) only for n 2 4, etc. Accordingly, one has the shells
36,&, 4f, 5d, Sf, ;1g, 6 4 Sf, 69, fih ptc. Since among the first 98 elemmts of

Table 3.3: Electron configuration of rare earths and actinides. In the right column
the respective closed shells are indicated.

Actinides
232 Chaptcr 3. Electronic slructure of semiconductor cr,ystals with perturbations

the periodic tablc, however, the 5g-shell already remains unoccupied, only
d- and f-shells are to be considered, namely the d-shells 3 d , 44 5 d , 6 d , and
the f-shells 4f and 5 f . The filling of the 3d-, 4d- and 5d-shells takes place
in the series of transztzon metals (together with the Elling of the 4s-, 5s- and
6s-shells). Among the transition metals, one distinguishrs the iron groixp in
which the 3d- and 3s-shells are being filled, the palladium group in which
the samc happens with the 4d- and 4.s-shells, and the platinum group where
the 5d- and 5s-shells are being Elled. The 4f-shells are being filled in the
rare earth elements, and the 5f-shells in the actinides. In comparison with
the s- and p-orbitals, the d - and f-orbitals have a smaller extension in spacc,
they lie mostly within the s- and p-shells of the same main quantum number
'l'herefore, mainly s- and p-electrons are involved in chemical bonding.
7 ~ .

This explains the remarkable chemical similarity of thc rare earth elements
with each other, a n d a certain similarity of these elements with the elements
of the main groups.

Complexes of point perturbations; associatos

Just as atoms arc bound in molecules when it is energeticaqy advantageous,


point perturbations also associate if the result is a state with lower energy.
They are called poznt perturbatzoa complezeu or u8souutP8. Bonding can
occur between various point perturbations: between chemically identical or
different impurity atoms, between impurity atoms and point defectu, and
among point defects themsclvcs. The associates may consist of two or of
several constituents. The diversity of these associates is comparably large
to that of molcculcs in chemistry. We give some examples below.

Donor-acceptor pairs

In the case of an ionized donor and an ionized acceptor, the lowering of


total energy through the formation of a bound complex of associtlteb is par-
ticularly obvious the two point perturbations are differently charged and
attract each other through electrostatic forces. This leads to the forma-
tion of donor-acceptor paws, in which the donor and acceptor atoms occupy
neighboring sites in the crystal. In general, the pairs are stable at several
possible distances, whirh gives rise to a variety of different donor-acceptor
pair complexes.

Di- and multi-vacancies

When there are two vacancies, the mechanism for the formation of bound
pairs can also be easily undertitood the (internal) surfacc of thc crystal i s
reduced if two previously isolated vacancies move together to occupy neigh-
3.1. Atomic structure of red semiconductor crystah 233

boring cryYtal sites. One caIls this associate a ddvacancy. Analogous atate-
ments holds for the association of more than two vacancies, which are called
multrvarunczes.

Frenkel defects

If, in a crystal, an atom moves froin a regular site to an interstitial site, then
it h v e x behind a vacancy which attracts the interstitial. Thus a defect pair
is formed in this process which consists of a self-interstitial and a vacancy.
It is called a Freibel defprt.
There arr important puint perturbation complexes which occur only in
a specific material or matrrial group. We now consider some examples for
si snd G A S .

Point perturbation complexes in Si

Wr various reasons, hydrogen is often present in Si. In p-type Si, H atoms


undergo chemical bonding with the availablc wreptoi atoms. Thereby the
electron of the H atom is captured by the acceptor atom: which becomes
singly negatively charged. One may also conclude that the acceptor expends
i t h hound hole to the H atom ralher than to the valence band because it has
lower energy there. In B-doped Si, for example, it negatively charged B-ion
and ti positively charged H-ion are formed in this way. The two ions attract
each other by- Coulomb forces, which results in the formation of a neutral
( H . B )pair. Of course, the pair will not be able t o arcept an clcrtron which
means that the B atom has lost its ability to act as an acceptor.
Chalcogen atoms like S, Se, and Te are incorporated in S i not only as
single atoms, but tlBo as two-atom molecules. Oxygen in Si enters into
bonding with a vacancy, forming a pair which probably constitutes the so
called A-renter known from caparity 111ea~iremmtfi.Oxygpn is also involved
in a wries of other defect complexes in Si, among others the so-called thermal
d a m r s . which are thusly named hPraiise of their origin in thwmal treatment.

Point perturbation complexes in G a A s

A prominent defect associate in GaAs is the so-called RX-center, which


acts as a donor. It is found in GaAs and also in (Ga, A1)As mixed crystals
under appropriate conditions (e.g., high pressure, for more see section 3.5).
Originally, the DX-center was attributed to a donor atom, like S&, bound
t o another point perturbation whose nature was unhiowii at t!hHt time and,
therefow, was denoted by X. Currently, the DX-center is thought to be due
to the donor atom alone, more strictly; to a donor atom which is incorpo-
rated interstitially but not substitutionally, as commonly happens. Another
234 Chapter 3. Hectronic structure of semiconductor crystals with perturhations

typical defect associate in GaAs and other Ill-V semiconductors is the com-
plex formed by an As-antisite defect (a Ga atom on an &-site) associated
with an As-interstitial. Now, this complex is believed to form t h e care of the
so-called b;L!Lcentei- in GaAs. This center manifests itself as deep level hav-
ing strong influence OR the electrical and optical properties of GaAs (more
detail on the EL2-rcnter may br found in section 3.5).

Point perturbation aggregates

I the complexes formed by point perturbations bmome larger and reach


mesoscopic size,one refers to them as aggw.qates. Complexes of macroscopic
size such as, for example, oxygen or heavy metals in Si cryshls, are called
precipitates.

Latticc relaxation

The forces on atoms in the irnmediatc vicinity of a point perturbation dif-


fer frorri those in lhr ideal crystal. They are non-xr?rol in general, at. the
ideal crystal sites. Thus the atoms are forced t.o move t.u new equilibrium
sitcs. This is known as lattice relamtion (in Figure 3.1 this effect is omit-
ted). The new rqiiilihriiim sites are initially tinknown. In principle, they
can be determined by means uf atomic structure calculatious for t h e per-
turbed crystal. These have t o be performed sirnultanmusly with calculations
of the elect,ronic structure, just as is done in self-consistmi. calcidations of
the electronic and atomic st.ructwes of ideal crystals described in Chapter
2, section 2.2. HoweveT, there is an impwtant difference between the two
cases. For ideal crystals, the calculation of at,omic structure may be avoided
since, for t,he latter, complete and reliable experimental data are available.
However, in regard t,o the atomic structure of a crystal in the vicinity of a
point perturbdion, in many cases, hardly more t,hari t.he symmebry is known
from experiment. Thus the self-consktent calculation of the electronic and
at.omir:sbriictures must actually be carried out. if lattice relaxat,ion becomes
important.
Experimental information concerning the symmetry of lattice relaxation
may be derivrd from observation of t,he Jahn-TelEev efleet. This eflect results
in splitting of the degenerate energy levels of a point perturbation due to
a symmetry-lowcring latt.ice relaxation. In the case of a varaiicy: for exam-
ple, the lattice rehxation can reduce the original tetrahedral symmetry T d
to tetragonal symmetv D2d. Such spontaneous symmetry breaking occurs
when it leads to lowering of the total energy of the crystal. This mag; happen
when, in the unrelaxed state, i.he point perturbation has a c1rgenerat.e level
which is only parhially occupied. First. of all, the degenerate level will split off
due to the symmetry lowering relaxation. According to perturbation theory,
3.1. Atomic structure of real semiconductor crystals 235

this splitting proceeds such that the center of gravity of the levels remains
unchanged. Thus, along with levels shifted up, there are also levels which
are shifted down. If only the latter are predominantly occupied, then the
energy of the electrons localized at the point perturbation decreases. This
energy reduction can compensate the increase in total energy due to the
removal of atoms from their equilibrium sites. If this happens the relaxation
is energetically favorable and will take place spontaneously.
In the case of the Jahn-Teller effect, the displacements of atoms are of
the order of magnitude of one tenth of an Angstrom, i.e. they are relatively
small. Larger displacements, of the order of magnitude of one Angstrom,
are observed at point perturbations for which different atomic structures are
stable, depending on external conditions as, for example, the position of the
Fermi level. This phenomenon is observed at the D X centers in GaAs and
(Ga, A1)As mentioned above (see section 3.5 for more detail).

3.1.3 Formation of point perturbations and their movement


Formation of structural defects

All of the above mentioned defects of ideal crystal structure may in fact
exist in real semiconductors. There are various reasons for their occurrence,
the most important and general being the second law of thermodynamics.
According to this law, the thcrrnodynamic cqiiilibrium state of a crystal at
temperature T and pressure p , is characterized by a minimum of the Gibbs
free energy G = H - TS. Here H is thc enthalpy and S the entropy of the
system. The entropy of a macroscopic state is proportional to the logarithm
of its microscopic realization probability (or thermodynamic probability)
and the proportionality factor is given by Boltzmanns constant k. The
system considered here is the totality of the atoms of the crystal. Lets
assume that there arc only chemically correct atoms, and that these are
randomly distributcd in space. Then the idedl crystal is formed wherein the
atoms move to the regular crystal sites. This corresponds to a very special
state of the system. It is extremely improbable compared to the large number
of states in which deviations from the ideal configuration of atoms appear
as described above. The entropy S of the ideel crystalline state i s smaller,
therefore, than that of the imperfect crystalline states.
The minimum of enthalpy II i s reached in the ideal crystalline state.
Since the entropy S takes larger values in other states, the minimum of H
is not necessarily coincident with a minimum of the Gibhs free energy G .
Depending on temperature, a minimum of G = H - T S can also be adjusted
for a state of the crystal which is less advantageous with respect to enthalpy,
but more advantageous with respect to entropy, i.e. for a b t a k with a non-
vanishing concentration of structural defects. As a concrete examplcs we
236 Chapter 3. Efectronic structure of semiconductor crystals with perturbations

consider a vacancy in an elemental semiconductor.


In an ideal crystal. the periodicity region contains a number J of identical
atorris. A vacancy is created w h m one of t h e atoms is removed from the
crystal. Altogether, there are J different possibilities for such a removal,
as many as thme are atorns. For each realization of the irlcal crystal one
has, therefore, J realizations of the crystal with B vacancy. II n independent
vacancies exist, the number o f realizations is . J ( J - I ) . . ( J - r L + l)/nL The
gain of entropy AS compared to the ideal cryatd t hrrefore mntnints to

As=kln[ J!
n!(J - It)!
1,
In a more rigorous treatment, the entropy of lattice oscillations has yet to
bP considered in AS, more accurately, the change of this entropy because of
the alteration of the spectrum of lattice oscillations in the presence oI the
vacancies. We will neglect this effect in what follows. What must, however,
he taken into account is the enthalpy Rf necessary for the formation of n
vacancy (at constant pressure). For R independent vacancies the enthalpy
requirement is n H f . Altogether, the gain of Gibbs free energy AG due to
the formation of n. vacancies becomes

This expression Ims t o be minimized with respect to TL for a h e d value of T


and p. According to of Stirling's formula, h { n ! )N n In(n) - n , the minimum
coiidition may be written as

Since n << d can be assumed, the ilctivalion 1.m~

follows from this relation. The enthalpy of formatiult HJ or B vacancy typ-


i d l y amounts to several c l ' . Assuming H f = 2 5 el/ and T - 1400 K
(roughly the crystalhdation temperature of Si), one gels the value n =
J x 5 x 10-l' from expression (3.5). Because J - s x cm-' (for Si),
a varancy conrmtration of about 1013 crriL3 follows. Similar concentrations
are ohtnincd for other point deferts. Thew v a h s refer. hy thcir derivation,
to the high temperatures assumed above, at which the c q s t d is grown.
Cooling down to room temperature often does not, however, substantially
change the defect concentrations The defects are frozrii in. From this one
may conclude that the presence of a considerable number of point defects in
3.1. Atomic struclore of real semiconductor crystals 237

crystals is essentially unavoidable. The Si singlecrystal bars used in micro-


electronics its a material base, in fact, contain vacancies and interstitials in
the above estimated concentration of about ~rn-~.

Incorporation of impiirity a t o m s

Chemical point perturbations are not as unavoidable as structural defects.


Of course, the laws of thermodynamics in this case also act in a direction
which leads away fro111 the absolute chemical purity of the ideal crystal.
This tendency, however, can only be effective to the extent that chemically
'wrong' atoms are available in the raw materials and the chamber in which
the crystal is grown. The state of a mure or less homogeneous distribution
of impurity atoms in the whole chamber, i.e. the state in which the growing
crystal also contains impurity atoms, has smaller Gibbs free energy than
that of the chemically absolutely pure crystal with all the impurity atoms
confined to the remainder of the chamber. However, by cleaning the raw
materials and the growth chamber in thermodynamic terms by an extrac-
~

tion of entropy - the number of chemically 'wrong' atoms can be reduced.


Theoretically, no limit exists for the degree of purity achievable, practically
such td limit is set by the cleaning expenses, of course. Therefore, one re-
duces the concentration of impurity atoms only to a level which is absolutely
necessary for the application iinder consideration.
A boundary for the achievable concent,ration of impurity atoms in a crys-
tal exists, as a rule, in the form of an upper limit. Under thermodynamic
equilibriiim conditions it is given by the solubility of the corresponding el-
ements in the semiconductor crystal. The solubility increases with rising
temperature acrording to a law which is similar to that for the vacancy con-
centration of equation (3.5), providcd H f is underst,ood as the formation
enthalpy of the impurity. The latter depends on the chemical nature of the
impurity atom and host crystal, and it also differs for incorporations at differ-
ciit crystal sites like substitutional or interstitial ones. High solubility values
are achieved, as a rule, if the element of lhe impurity atom is chemically
similar to at least one of the elements of the crystal. Incorporation at sub-
stitutional sites is preferred in this case. If the crystal consists of elements of
tho main groups, which i s the case for the majority of known semiconductor
materials, one has relatively high solubility, especially for impurity atoms of
these groups themselves. In the extreme rase, an alloy will be formed with
t,he impurity atoms. The dissolving of A1 in GaAs, for example, results in
a (Gtt, Al)As alloy. In such cases, t,he solubility equals the concentration of
the host crystal atoms, i.e. crn-'. For P in Si or Ge in G A S , solubility
values of lo2' c n ~ -are ~ reached, which lie just a little below. For transi-
tion metal elements, which d i k r st,rongly chemically from the main group
elements, t>hesolubilities in semiconductors made of such elements are sub-
238 Chapter 3. Electronic structure of semiconductor crystals with perturbations

stantially smaller. They lie in the region between lOI4 cm-3 in the case
of the elemental semiconductors, where the transition metal atoms can be
incorporated both substitutionally and interstitially, and 1017 ~ r n -in~ the
case of thc compound semiconductors where, as a rule, substitutional incor-
poration is preferred. Generally, solubilities for interstitial impurities depend
strongly on the sizes of the impurity atoms. Small atoms, like H, may be
interstitially incorporated in higher concentrations than larger atoms, such
as atoms of the transition metals. It is remarkable that rare earth atoms are
predominantly incorporated substitutionally in tetrahedral semiconductors
of the main group elements. This may be understood by means of the fact
that the not-yet-completed 4f-shells are smaller than the already filled 5p-
and 6s-shells, so that 3- and pshells are involved in chemical bonding, just
as in crystals of main group elements. The solubilities of the rare earths are
nevertheless relatively small, typical values being of the order of magnitude
I 014 cm-3.
There are various procedures to introduce impurity atoms into crystals.
The easiest one exploits the growth of the crystal from the melt: the impuri-
ties are added to the melt of the host material, &henthe melt is rooled down
to cause crystallization. Another method proceeds in the solid state: the irn-
purity atoms are diffused-in from an outer source. To speed up diffusion,
the crystal is heated. Besides the two processes mentioned, which introduce
impurity atoms under equilibrium conditions, there are also non-equilibrium
procedures to reach this goal, among them the so-called i o n i m p l a n t a t z o n
In the latter, the impurity atoms, after they have initially been ionized and
then accelerated by an electric field, penetrate into the crystal where they
are implanted. This implantation process must be followed by heating of
the crystal in order to heal out the defects which arise during implantation
(annealzng). The heating also makes it possible for the impurity atoms,
which were initially placed at sites relatively indiscriminately, to reach their
equilibrium sites. If the concentration of implanted impurity atoms should
be larger than the equilibrium solubility, the surplus atoms are later ex-
cluded from the crystal to a certain extent (precipitation). In many cases, a
relatively large number of surplus atoms remains in the crystal, which then
is in a frozen-in non-equilibrium state.

Migration a n d diffusion of point perturbations

Point perturbations interact with the atoms of the crystal surrounding them.
Both the point perturbations themselves as well as the environment atoms
are in permanent motion because of the lattice oscillations. This results in
more or less chaotic forces on the point perturbations. Due to the effects
of these forces the perturbations move over to equivalent crystal sites in
adjacent primitive unit cells. This is referred to as a m i g r a t i o n of point
3.1. Atomic structure of real semiconductor crystals 239

perturbations. The migration of a vacancy proceeds such that an equivalent


neighboring atom moves to the vacancy site and leaves a vacancy behind
at its former position. The latter vacancy is displaced with respect to the
original vacancy. A substitutional impurity atom migrates mainly with the
aid of a vacancy, namdy by filling a neighboring vacancy site. In contrast
to this, an interstitial impurity atom does not need the help of a vacancy for
its migration, it can move directly to an adjacent equivalent crystal site.
The initial and final states of an elementary migration act are relatively
stable states of the crystal. Between them, intermediate states with larger
total energies occur. This means that ttn energy barrim has to be overcome
in an elementary migration act. One calls it the magration barrier Em. In
order for migration to occur, the lattice oscillations must provide the energy
Em or the enthalpy Hm = Em + p V , with the pressure p held constant
rather than the volume V , as comnionly OCCUTS. In this way, the migration
rate becomes proportional to the activation factor e x p ( - H , / t T f . For the
migration af a vacancy in Si and Ge, H, amounts l a about 1e V , and for the
migration of a self-interstitial in these materials, it is smaller than 0.25 el/.
Migration constitutes the elementary microscopic event underlying the
mtacroscopic process of diffusion. The latter occurs then when the distribu-
tion of the migrating point perturbations is spatially inhomogeneous. For
not-too-large gradients of the defect concentrations, the first Fick's law may
be applied in many cases. Conespondmgly, the diffusion of a particular
kind of point perturbation may be traced back to just one material con-
stant, namely the diffusion coefficient D. The temperature dppendence of
the latter also exhibits an activation behavior.

where Do is the limit of U at high temperatures. In the case of diffusion


through interstitial sites, the activation enthalpy Q equals the correspond-
ing migration enthalpy H,. For diffusion through vacancies, Q is given by
the sum of the migration mthalpy Hm and the formation enthalpy H f for a
vacancy. Substitutional impurity atoms preferentially diffuse through vacan-
cies. In Table 3.4 the Q-values for substitutional and interstitial diffusion of
some impurity atoms in Si are listed. For P in Si, which diffuses substitution-
ally, the diffusion coeficient at 1500 K is about lo-'' cm2 sec-'. At1 atoms
in Si preferentially diffuse through interstitial sites. The d i f f ~ scoefficient
i~~
at 1500 K of about 2 x c m 2 scc-l exceeds that for P in Si by 6 orders of
mappitude. The reason for this hiigp d i k e n c e is thp siibstantially smaller
activation enthalpy for interstitial diffusion as compand to substitutional.
240 Chapter 3. Electronic structure of semiconductor crystals with perturbations

impurity atom B Ga P H Au Cu
activation enthalpy (eV) 3.7 3.9 3.7 0.5 0.39 0.43

3.1.4 Line and planar defects

With an increase in the number of point perturbations, a tendency arises


for these perturbations to arrange themselves on lines or planes of macro-
scopic size. Then one has 1- and, respectively, 2-dimensional perturbations
in contrast to the 0-dimensional ones considered above. In practice, order-
ing on lines and planes occurs only for structural point perturbations. The
most important representatives of a 1-dimensional perturbation are step-
and screw dislocations. At a step dislocation, the occupation of a particular
lattice plane breaks down along a line of lattice points - the plane is occupied
on one side of this line (representing the step dislocation line), but not on
the other side (see Figure 3.2). If one calculates the line integral over the
path shown in Figure 3.2 (left-hand side), then the result will not be equal to
zero, as would be the case for an ideal crystal, but equal to a non-zero vector
perpendicular to the dislocation line. It is called the Burgers vector. For a
screw dislocation one has to view the crystal as being cut by a semi-infinite
lattice plane, which is bounded by a line of lattice points representing the
screw line. Then the lattice planes left and right of the cutting plane are
shifted parallel to the screw dislocation line by a lattice vector. After that
one reconnects the two crystal halves again (see Figure 3.2, right-hand side).
The line integral about the screw dislocation line yields a vector parallel to
this line. In a dislocation, deviations from crystal structure are, indeed, not
limited to a line as one could initially assume. Slight atomic displacements
(strain) occur in a finite macroscopic environment. The microscopic core of
the perturbation is limited, however, to the dislocation line.
The two most important examples of 2-dimensional perturbations are
stacking faults and grain boundaries. In a crystal with stacking faults, the
various lattice planes carrying the atoms are not stacked in the same way as
in the ideal crystal, but certain lattice planes are twisted by an angle. Grain
boundaries occur in crystals which in fact consist of two differently oriented
half crystals - the grain boundary is the lattice plane at which these two
half crystals meet. In a sense, surfaces of crystals may also be considered as
2-dimensional perturbations.
3.2. One-electron Schriidinger equation for point pcrturbations 24 1

Figure 3.2: Illustration of a step dislocation {left-handside) and a screw dislocation


(right-hand side). ( A j L w f35w.. 1990.)

Having surveyed of the atomic structure of real semiconductor crystals we


will now proceed to the electronic structure of such crystals. In this, we
restrict, our considerations to crystals with compositional or structural point
perturbations, therefore, we excliide 1- and 2-dimensional perturbations.
'rhis is done in recognition of the fact that only point perturbations can play
an important positive role in semiconductor devices, while perturbations of
higher dimensions are generally disruptive. Fortunately, the latter can also
be more easily avoided than 0-dimensional perturbations, because they are
less effective in increasing entropy. The silicon bars used in microelectronics
are now grown practically free of dislocations and grain boundaries.
In the following sect'ion we will formulat'e the Schrijdinger equation for
the oneelectron states of a crystal with a point perturbation.

3.2 One-electron SchrGdinger equation for point


perturbations
In regard to its geiwral form, thr ontLclectrou Schrodinger q u a t i o n for a
crystal with point perturbations does not substantially differ from the o n e
electron Schriidinger equation for an ideal crystal. The reason for this is that
the derivation of the latter equation in section 2.2 never actually involved
the periodicity of the ideal lattice. Only the oue-elwtron potentid limrt(x)
of the peTtriirbeclrrystal diffws from the rarresponding potential V ( x )of the
ideal crystal given in equation (2.76). However, just like the latter, also the
potential V ~ " ( x of ) the perturbed crystal is the sum of three contributions,
the potential V ' c ? 7 L ( ~ ) caused by thf atomic cores of the perturbed crystal,
the Hartret. yotentid I/Frt(x) of the valence electrons, and the exchange-
correlation potential V F
:(
x
) of these electrons, In analogy to equation
(2.76) we therefore have
242 Chapter 3. Electronic structure of semiconductor crystals with perturbations

VWTt(X) = VL?'t(X) f V,"'"x) + Vg;t(X). (3.7)


With this, thc one-electron Schrodinger equation for the wave function $ l V ( x )
of an electron of the perturbed crystal reads

3.2.1 Elect ron-core interaction


First of all, we discuss the potential VcPTt(x)
of the electron-core interaction.
We decompose it into a sum of the core potential Vc(x) for the ideal crystal,
and a potential V,'(x)which describes the change of the core potential due
to the point perturbation. Thus,

VC~'"X) = K ( X ) I V,(x). (3.9)


In further clisrussion, we assume that only one point perturbation exists in
the crystal. The center of this perturbation will be taken as the origin of our
Cartesian coordinatc systcm, and we procecd on the assumption that the
potential VJx) falls off with increasing distance form the origin and finally
approaches zero:

V,'(x) -+ 0 for Ix I + 00. (3.10)


The point perturbations listed in section 3.1 differ as to how fast this decay
proceeds.
We consider, initially, the case of a substitutional impurity atom whose
core contains almost the same numbers of protons and neutrons as the core
of the host atom. With this rqiiirement, the two atomic cores differ mainly
through their different charges. One refers to them as isocoric impurity
atoms. This case occurs, for example, if a P atom with only one additional
proton and one addit,ional neutron substitutes a host atom of a Si-crystal (as
illustrated in Figure 3.3). We denote the number of (positive) elementary
charges of the core of the impurity atom by Z r , and the number of (positive)
elementary charges of the core of the host atom by Z H . Because of charge
neutrality of the individual atoms, ZI and ZH are simultaneously also the
numbers of valence electrons of these atoms. The potential energy of a
valence electron of an impurity atom of the type described differs, above all,
by the change of Coulomb potential from the potential energy of an electron
of the host atom. If one considers the cores to be point-like and neglects
spatial dispersion of dielectric screening, then one has, approximately,
3.2. Oneelectron Schrcdinger equation for pint perturbations 243

(3.11)

where E is the static dielectric constant of the semiconductor material. Devi-


ations from this perturbation potential are to be expected in close proximity
to the impurity atom, on the one hand, bwansc there the spatial dispersion
of dielwtric screening cannot be neglected, and. on the other hand, because
the core of thp impurity atom diffcrs not only by its charge number from
that of the host atom, but also in other respects. In fact each core has a
iinite spatial extension because of its spatially extended r o w electrons (the
nucleus of the core may be treated as a point charge). In close proximity
to the center, the core electrons give rise to additional forces, heside t h e
elerlrostatic point charge forces already counted in [3.11). These additional
forces are caused by higher moments of the core electron charge distribution,
and also by exchange and correlation effects. If the cores of the host and
impurity atoms differ, whch always happens iI the atoms arc not identical,
the additional core forces will also differ. Both effects, the spatial dispersion
of screening and the differpiirps of additioiiat rorc forces, are jointly termed
central cell correcfzonu.
The perturbation potential (3.11) is distinguished in that. with the ex-
clusion of a close environment of the impurity atom, its variation over a
primitive unit cell is relatively small. which implies that it remains signifi-
cant over a relatively large distances from the impurity atom compared to the
lattice constant. In this context, it is called a long-range potential. We may
state. therefore, that isocoric impurity atoms are approximately described
by smooth or long-range perturbation potentials.
The potentials which apply for non-isocoric impurity atoms are different.
Consider first the case in which the charge eZf of the impurity atom core
equals the charge eZH of the host atom core. This does not necessarily
mean that the nuclei of the two atoms must have the same numbers of
core electrons, because different numbers of protons of the atomic nuclei
may compensate the difference of core electron charge. The requirement of
equal numbers of core charges implies, however, that the numbers of valence
electrons of the two atoms must coincide - as described by the term asovalent
impuritr atoms. Examples for non-isocoric isovalent impurity atoms are
C atoms in Si- or Ge-crystals, or N atoms substituting P atoms in Gap.
The potential energy of an electron at a non-isocoric. isovalent impurity
atom differs from the potential energy at the host atom not just by the
screened Coulomb potential (3.10). but by the different core electron shells
in the two cases. The perturbation potential b:(x) accordingly contains only
electrostatic contributions of higher moments of the core electron charge
distribution difference, as well as exchange and correlation contributions
due to this difference. Both kinds of contributions decay more rapidly with
244 Chapter 3. Electronic structure of semiconductor crystals with perturbations

- _ _ _ Si_ Si_ _
Si Si Si Si Si

v; = v, pert -v,

v
Figure 3.3: Illustration of thc origin of long-range (left-hand side) and short-range
(right-haritl side) core perturbation potentials.

increasing distance fiorn the center than does tlic Coulomb potential of a
point charge, they are therefore described as short runqe. These arc also the
potential contributions, whicb in the context of the ibocoric impurity atoms
abovc, gave rise to central cell corrections.
The exact &termination of the pertrirhatiori potential is a problem which,
like the determination of the periodic core potential of an ideal crystal, can
ultimately be solved only by numerical calculations. These show, in fact,
that the pet turbation potentials of isovdlenl impurity atoms differ horn zero
substantially only over a distance of a few lattice constants. Consequently,
we have

v'(x) 0 for I x I 5 a f e u lattice constants


v(;(x)
(0 for all other Ix1
Short-range perturbation potmtials apply not only to isovelcnt, non-isocoiic
(3.12)

substitutional impurities, but also to structural defects such a8 vacancies


or interstitials. l n thc case of a vacancy, the occurrence of a short-range
perturbation potential is particiilatly obvioiis. The remuval o f a Si at<nrn
from the chain of S i atoms in Figure 3.3a yields the potential profile of
VJx) +V,'(x) as depicted in Figure 3.3b, and the perturbation potential
has the form shown in Figure 3 . 3 ~ The
. latter approximately represents the
ncgative potential v,'(x) of the missing atom.
3.2. One-electron Schrljdinges equation for point perturbations 245

Finally we foiego the requirement that the core charges of the two non-
isocoric atoms be the same. Examples of this most general case are Cd
atoms with their 2-fold core charge in a crystal of S i atoms with their 4-
fold core charge, or Sn atoms with their Cfold core charge on a G&site
in GaAs, thereby substituting a triply positively charged host atom core.
In these cases, the perturbation potential represents a superpobition of the
screened Coulomb potential (3.11), which accounts for the different core
charges, and a short-range potential which takes accoimt of all remaining
differences bctween the cores d5 well as aspwts of the spatial dispersion
of screening beyond those accounted for in equation (3.11). The effects
of the two potential contributions are not independent of each other. The
same bhort-range potential has a greater effect if the Coulomb potential, due
to a larger core charge difference between the impurity and host atoms, is
stronger. This comes about because the stronger Coulomb potential pulls the
electrons closer to the core, where the the short range potential is essential.
Just like the elwtron-core inteiaction potential, the Hartree potential
V,(x) and the exchangecorrelation potential V,,(x) also undergo chan-
ges in the prrsence of point perturbationb. Below, we describe these changes
for the Hartrre potential V I y t ( ( x )and the exchange potential V,((x) of
the liartree-Fock approximation. Similar results can be derived for the
exc~iaiig~rorrt~ation potential vzt(x)of the LDA mettiod.

3.2.2 Elect ron-electron interaction

Hartree potential

The Hartree potential V r t ( x z ) of the r-th electron is, as before (see relation
2.49), given by the expression

(3.13)

However, the oneparticle states $,(x2) here are not those of the ideal crystal,
but those of the pertiirbed crystal and the summation runs over the o n e
particle states u i occupied in the ground state of the pcrturbwl crystal.
From the physical point of view it is quite clear that the prrtiirbed rrys-
tal ~ H two
S kinds of stationary one-particle states; first, those with energy
eigenvalues which were already allowed for the ideal crystal, i.e. those within
the bands and, secondly, those with F I I C I eigenvalues
~~ within the energy gap
between the valence and conduction bands. The states of the first kind rep-
resent pure Bloch states only in zero-order approximation with respect to
the perturbation potential, while, in higher approximations, superpositions
246 Chapter 3. Electronic structure of semiconductor crystals with perturbations

of other Bloch states have to be added due to scattering by the perturba-


tion potential. The pure Bloch state property of being infinitely extended
over the entire crystal also applies to these scattered states for most of the
energies. The states of the second kind, which belong to energies in the for-
bidden zone of the energy spectrum of the ideal crystal, are localized, unlike
the Bloch states, at the position of the point perturbation (formal proofs of
this will be given in sections 3.4 and 3.5). We identify the extended states by
quantum numbers v " ~ ,and the localized states by quantum numbers doc.
Furthermore we assume that, in the ground state, n of the total number N
of electrons of the perturbed crystal are found in localized states v;OC O, and
the remaining N - n electrons are in extended states viztd '. Accordingly,
we decompose the Hartree potential of the perturbed crystal into two parts,
the part V Z t d ( x i )stemming from electrons in extended states vrtd O, and
the part VAOC(xi)due to electrons in localized states v;OC',

(3.14)
with
extd
V g t d ( x i )= e2 / d3x'
k J
and

(3.16)

Here we assume that the electron i on which the potential acts is located at
the center. This assumption will also be maintained below.
The localized part VhOC(xi)of the Hartree potential depends on the num-
ber n of electrons at the center. There are two sources of this dependence.
First, the number of electrons which enter changes with n, Vfi"(xi)becomes
more repulsive if n is large, and less so if n is small. Second, the wave
'
functions of localized occupied states vLOC over which the sum in (3.16) is
extended, depend on n. They become more localized if electrons are removed
from the center, and less localized if electrons are added. One says that the
wavefunctions relax.
For reasons similar to those causing the localized part VhOC(xi)of the
Hartree potential of the perturbed crystal to depend on the number n of
electrons at the center, the extended part V;&(xi) also differs from that
of the ideal crystal, firstly, because the number N of extended electrons
is decreased by the localized ones and, secondly, because the states of the
extended electrons relax. The first change causes a relative potential cor-
rection of order of magnitude 1 / N and thus can be omitted (a comparable
3.2. One-electron Schriidinger eqmtim for point perturbations 247

approximation was employed previously in the derivations of the Hartree


potential and of Koopmans theorem in section 2.2). The change of the
extended states is essential, however: because all N - n occupied states are
affected. In this connection, the probability amplitude of the extended states
is reduced close to the localized electrons of the center because of Coulomb
repulsion. Hence, a positive excess charge arises in the vicinity of the local-
ized electrons which screens the Coulomb potential of these electrons. This
change may be accounted for by replacing the localized part VAm(xi)of the
perturbed Hartree potential by a screened potential Vh(%), and simulta-
neously substituting the extended part V F d ( x i )of the perturbed Hartree
) the ideal crystal. Accordingly,
potential by the Hartree potential V ~ ( x iof
we set Vkm(-) --t VA<xi)and V r t d ( q ).+ V~(xi). Then it follows

(3.17)

where, by definition, VA(xi) is given by the expression

where e - ( q , x) is the inhomogeneous position-dependent screening func-


tion of the crystaL

Exchange potential

As we know from section 2.2, the exchange potential describes the Coulomb
interaction with the exchange hole which occurs because two electrons of the
same spin cannot reside at the same position. For the localized electrons of
the perturbed crystal, the positional uncertainty is smaller than that of the
extended ones. Accordingly, their exchange interaction will be stronger than
that of the extended electrons. It is therefore again expedient to decompose
the entire exchange potential V,Pt(x,),which an electron a localized at the
center feels, into the two parts V,..td(&) and l r ~ ( x of,
~ )respectively, the
(N - n ) extended and n localized electrons. We replace the extended part
by the exchange potential V x ( x i ) of the ideal crystal and simultaneously
substitute the localized part by an effective exchange potential Vi(xz).With
this replacement, we have

= VX(Xi)
v,pt(X,f + Vfi(X2). (3.19)

Using relation (2.61), it follows that


248 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.20)
where the summation over k inrlirdes only particles whose spin W k equals the
spin LT? of particle i.
Substituting trhe three potential parts (3.91, (3.17) and (3.19) into the
oneelcctron Schrbdinger equation of the perturbed crystal, the various terms
may bc arranged such that the effective oneelectron potential E'(x} of the
ideal crystal (2.60)occurs. The equation reads

The three perturbativc potential parts in this equation have different ef


fwts. In gcnpral, the core perturbation potential l<'(xz)is the strangest.
The occiirrenre of localized states at the center. the central feature of our
considcrations above, is maidy due to this potential contribution. If only one
electron is localird at the center, tlip two othw potential parts L$(x%)and
Vdk(x,)due to electron-electron interactions vanish completely. <;enerafly,
they a e non zero and lead to corrections of the localized eigensolutions of
ihe Schriictinger PqiiBtion (3.21) which rontaiiis L:(x) as the only potential.
'I'hcse corrections will now be estimated by means of perturbation theory.

Pert rirbation theory

The Srhrodingrr equation of zcro th order reads

[ H f q x ) ] ls'v(x) - E U T h ( X ) , (3.22)

+
whcrt. I1 - p2/2na V(x), as bcfore. signifies the oneelectron Hamiltonian
operator of the ideal ciystal. The index i is omitted because all electrons
feel the same core perturbation potential. If only one particle is localized at
the center, equation (3.22) is exact.

Hartree energy

The energy correction due to the Hartrcc potential for an electron z in local-
ized stateqVt i s given in first order perturbation theory b - ~the rxpectation
valri~($V, 1 VAw I vv,).[!ring (3.181, this expression may be written as
3.2. One-electron Schrodinger equation for point perturbations 249

(3.23)
We extract the factor (E-'(x, x')/ I x-x' 1) from the x'-integral by evaluating
it at an average position X in place of x'. The remaining x'-integral yields 1
because of the normalization of the wavefunction @,,(x'). The integral over
X, multiplied by e 2 , will be denoted by Uy,, i.e. we set

(3.24)

With this, the expectation value of the perturbation of the Hartree potential
becomes

The energy Up' of ( 3 . 2 4 ) is called the Hubbard energy.

Exchange energy

Next we will calculate the corresponding energy correction due to the pertur-
bation of exchange potential V%{x)of (3.19). The latter depends on the spin
of the electron because. contrary to the ideal crystal case where the num-
bers of electrons with 'spin-up' and 'spin-down' are equal in ground state,
the corresponding numbers n~ and nl of electrons localized at the center can
differ. In terms of the total number n = n~+nl and the total spin projection
M S = (1/2)(nT - "1) of the localized electrons, the two partial numbers ny
and nl may be written in the form

1 1
"1 = - ( n
2
+
2 h l s ) , nl = -(n - 211iIS) .
2
(3.26)

For the expectation value (&,t 1 1 5 I iu,) of the perturbation of the exchange
potential, equation (3.20) yields the expression

I ~ I'+v,)
( ~ v , i -(no, - I ) J , f o r not 2 1. (3.27)
where g, is the spin quantum number of the state v,, and

(3.28)

is so-called the exchange integrcab The dependence of this integral on the


orbital state v' = Uk has been ignored in expression (3.28).
250 Chapter 3. Electronic structure of semiconductor crystals with perturbations

To solve the Schrodinger equation (3.21) explicitly) the number R of electrons


at the center must be known. This question will now be addressed.

Number of electrons at the center

To say that an electron is at the center means occupying a state which is


IocaIized there. The number n. of electrons at the center equals, therefore,
the number of electrons in localized states. This number is initially unknown.
It can only be determined self-consistenbly because the Hamihonian which
determines the states depends on it. In thermodynamic equilibrium and at.
temperature T = 0 K the number of electrons at the center may be found
by self-consistently counting the number of localized states which belong to
energy eigenvalues below the Fermi level. Since the position of the Fermi
level depends on the thermodynamic state of t,he semiconductor, the number
n. of electrons at. the center is also a state dependent quantity.
Below, we introduce some common not,ations used in this context and
estimate n for particular impurity atoms whose chemical bonding to the
host crystal is known. We assume the point perturbation to be elechically
neutral, which means that in generating it, the same number of positive and
negative elementary charges were added to the ideal crystal (as in the case
of an impurity atom on an interstitial site): or removed from it (as in the
case of a vacancy) : or removed and added (as in the case of a substitutional
impurity atom). Depending on i t s environment in the crystal, the paint
perturbation enters into a more 01 less strong chemical bonding with the
surroundiug atoms. A phosphorus atom in a Si-crystal, for example, is
chemically bound like a Si atom, i.e. the .Y- and p-states of the P atom are
involved in the formation of the valence band states of the crystal, and 4
of its 5 a2p3-valenceelectrons are hosted by these states. Therefore, only
1 of the 5 valence electrons of the P atom i s available to occupy b c a l i z d
states in the energy gap. This means that n = 1 holds for the neutral
phosphorus substitutional impurity in Si. If a Si atom in a Si-crystal is
substiluld by a boron atom, one has TI = - 1. Therefore, 1bole is available
to occupy the states localized at the center. Similarly, n -1 emerges if a P
7

atom in P ZnS-cryst81 replaces an S atom. The number V of electrons of an


impurity atom X involved in its bonding to the crystal and which, thus, have
energies in the valence barid and not in the gap, is generally referred to as the
oddation state of the atom. The latter is denoted by X + . A phosphorus
atom, for example: which is installed in a si- c r y s t d on a regular crystal-site,
has the oxidation state P4+, This notation, which looks like the number of
e1ernent.at-y charges at the atom without being such, originates from its prior
use for crystals and impurity atoms with pureIy ionic bonding. In the latter,
the valence electrons in fact pass from the impurity atom to the surrounding
crystal and the impurity atom is left behind as an Xv+-ion. This is no
3.2. One-electron Schrodinger equation for point perturbations 25 1

longer true if the bonding is covalent or partially covalent. Then some of the
bonding electrons remain at the impurity atom, and the true number of its
elementary charges differs from V. The n valence electrons of the impurity
atom, which remain after the departure of the V electrons into the bonds
with the crystal, and are available for occupation of the localized levels in the
energy gap, are sometimes called active electrons. In the case of an impurity
atom with Vi valence electrons one has n = Vi - V active electrons.
So far, the charge state of the impurity atom prior to its incorporation
into the crystal was taken to be neutral. However, positively or negatively
charged ions can also be introduced into the crystal, and atoms introduced
in a neutral charge state can have electrons removed or added within the
crystal. Similar statements hold for structural defects. If a particular per-
turbation center X is not neutral for the lack of Q electrons, in the sense
just specified, one says that the charge state of X is Q and writes X ( Q )
(where negative Q mean surplus electrons). The charge state should not be
confused with the oxidation state. In the purely ionic case, the distinction
between the two can be expressed most easily: the charge state counts the
elementary charges of the atom outside the crystal, and the oxidation state
counts its elementary charges inside. Generally, the oxidation state of a cen-
ter in charge state Q will be X('+Q)+ if the oxidation state of the neutral
+
center is V . The reason is that, in this case, V Q valence electrons are not
available for occupation of localized states. The number n of electrons which
are available, i.e. the number of active electrons, amounts to Vi - (V Q )+
if V,, as before, denotes the number of valence electrons of the neutral im-
purity atom. The notations for the oxidation state and the charge state are
summed up in the common symbol X('+Q)+(Q+).
A simply ionized sulphur atom in a Si-crystal, for example, is denoted by
S5+(l+).Of the 6 valence electrons of the S atom, 6 - 5 = 1 are available
for occupation of localized energy levels in the gap, instead of 2 electrons in
the case of the neutral center S4+(O+). Transition metal (TM) atoms can
be installed in Si- and other tetrahedral semiconductor crystals both sub-
stitutionally as well as interstitially. The oxidation states V and, therefore
also the numbers n of electrons at the impurity atom, are different in the
two cases. For substitutional TM atoms, V equals the number of electrons
which are left at the atom after it is bound to the crystal. Interstitial im-
purity atoms are only weakly bound, i.e. the number V of electrons of the
T M atom which occupy bonding valence states, is almost zero. The oxida-
tion state is therefore T M o f . Oxidation and charge states coincide in this
case, and the number n of electrons at a TM atom equals the number of its
valence electrons. Interstitial Fe atoms in Si are found in the oxidation state
F e o f , substitutional in the oxidation state Fe4+. In the first case, n = 8
(six electrons in 3d-orbitals and two in 4s-orbitals), and in the second case
the value of n = 4.
With the background analysis set forth above, we are now sufFiciently p r e
pared to address t h e solution of the one-electron Schriidinger equation for the
crystal with a point pcrturbation. Which solution methods can be applied
with succcss depends decisively on whethe1 lhe core yrrt iulmlion potential
is long- or short-range. Of course, t h r lattice translation symmetry, which
drastically simplifirs the bolution of t h ~Schriidinger eqiiation hi the casr
of an ideal crystal is perturbed by both kinds of potentials. but only for
short-rangepotentials in an e s s e a t d mannet-. For long-range potentials the
dec iations from lattice trauslation symmetry are relatively weak. In this
case the Scfirodinger equation, in a certain sense, may be decomposed into
two quations, one for the periodic potential of the ideal crystal, which was
solved previously, and one for the perturbation potential. T he latter is called
the LPffectivemass equation, which we will derive in the following section.

3.3 Effective mass equation


ure start with the Schrodinger equation (3.8) of the perturbed crystal in
the form (3.21). The most, iiriport.ant assumpt.ion we will make relat.es to
Ihe core part V ~ ( X of
) the pert,urhatiozt potential V A T t ( x )in this equation.
This potential part is supposed to be smooth on the atomic length scale. In
conjunction wit,h t,his wxmption, point pcrtiirhatioiis with shos-t-range core
pert.iirbation pot.entials Vf(x)are ruled out from the very beginning,because
these caiinot be considered to I c smooth. Point perturbations with Coihm-
bic core perturbation potentials are still allowed, although the Coilonib form
of this potential is not necessary for the derivation of t.he effective mass equa-
tion. The other parts of thr t o l d perturbation potential V&.t(xj of eqiiation
(3.21)!i.~. the pdurbations of the Hartree and exchange potentials l7L(x)
and ViCx), arc automat,ically smooth if V&(x) has this property because
these potential parts are dctcrmined by t.he solutions qVof Ychrodinger equa-
tion (3.21) which are smooth if L;(x) is also. The wwelunction dependence
of Vh(x) and Li(x) requires self-consistent solutions of the Schriidinger
equation. In this section we do not intend to solve this quabion explicitly
hut iclther to t,ransform it into another equation, the cffective mass equa-
tion? aliith can be solved more easily. Althoiigh the self-consistency demand
is transferred to the effective mass equation. it does not iulerfere with its
derivation. For the latter? Vh(x) and Vk(x) may be treated as smooth ex-
ternal potdialii which, together with V,(x)! add lip l o form a smooth total
perturbation potential i(x). Far hhe derivation which follows, the PO-
t.ential need not be t.he smooth perturbation potential VieVt(x)of a point
perturbelion, any smo0t.hpotential U ( x ) is allowed. This is important inns-
much as it becomes possible in this way to utilize the effective mass equalion
not only 01 point perturbations with Coulombir core perturbation poten-
3.3. Effective ma95 equation 253

tials, hilt also for macroscopic perturbations, such 8s that associatd with
an external electric Geld, which are smooth on the atomic length scale by
definition.
The goal of the following consideralions is l o simplify the oneelectron
Schriidingcr eyiiation

(3.29)

for t,he perturbed crystal in stages, so that., ultimately: a mare easily solvable
equation, namely the effective mass equation, results. In this mat,ter, we
will employ t,he fact that, in the vicinity of critical points of a certain band
v , the energy of an electron of an ideal crystal depends quadratically on
quasi-wavevector k, i.e. in the same way that the energy of a free electron
depends on its momentum. However, the free electron mass 7 n is replaced
by the effective mms rnc of the particular band and critical point chosen.
The e&ct.ivc mass includes the effects of interaction of the electron with the
periodic potential of crystal. Therefore, it is generally a tensor, and if il can
be reduced to a scalar maw, the value of the lat.ter generally differs horn the
mass of a free electron. Having in mind the effective mass description of the
band energy versus quasi-wavevector k relat,ion, one may suspect that the
influence of a perturbation potential V(X) on an &&on of the cryisla1 can
be calculated in an appr0ximat.e way as follows: One eliminates the periodic
potential of t h e oiic cltrtron Schrdinge~equation for the perturbed crystal!
while simultaneously replacing the free mass m in the kinetic energ); operator
by the effective mass m:. The resulting Schrodinger equation represents
the one-bend e#ective mass eyuatian in its simplest form It can be solved
much more easily than the original Schrodinger equation? which includes
the periodic crystal potential ill explicit form. For a siibstitutioiial P atom
in Si, for example. t.he effective mass equation is twsentinlly the same a8
the Schrodiuger equation for the hydrogen atom whose solutions are already
known.
The procedure described above needs, of COZITSC, further justificat.ion. To
provide this, quest,ions have to bc addressed which have been left, open so
far, for example, how the wavefunction of the e f l d i v e mass equation reiates
to the wavefunction of the original Schrodinger equation, and the matter of
what conditions must be placed on the perturbation potential V(x) and the
wavefunction 7;j(x)in order for bhe effectivp mass equation to be applicable.

3.3.1 Effective mass equation for a single band


To address these questions and derive the effective mass equation for a single
band, we rewrite the Schrodinger equdion (3.29) in the Bloch representation,
254 Chapter 3. Electronic structure of semiconductor crystals with perturbations

expanding ~ ( xwith
) respect to the complete set of Bloch functions pvk(x),
whence

(3.30)

with the expansion coefficients given by

(vk I = d3xlpXx)~(x). (3.31)

Since the (pYk(x) are eigenfunctions of H with eigenvalue E,(k), the Schrodinger
equation (3.29) takes the following form:

+
[ E y ( k ) 6 v f y f 6 k k ~ (vklUjvk)] (JklV) = E ( v k l $ ) . (3.32)
vk

To transform this equation into the effective mass equation, the assumption
made at the outset that U ( x ) should be a smooth potential, must be specified
further. This is done by assuming that the change of U ( x ) over a primitive
unit cell is small in comparison with the change of the periodic crystal po-
tential t ( ~over
) such a cell. To formulate the condition for smoothness in
this sense quantitatively, we decompose U(x) in a Fourier series, using the
same notation introduced previously in Chapter 2. We have

(3.33)

with Fourier coefficients

(3.34)

The sum in (3.33) is extended over the entire infinite k-space, meaning over
all Brillouin zones. Smooth functions U ( x ) in the above sense have Fourier
coefficients (klU) which, for large k-vectors, more strictly speaking, for k-
vectors outside of the first BZ,are small compared to the Fourier coefficients
(kll.) of the periodic crysta1 potential V ( x ) . The latter components were
calculated in section 2.4. According to formula (2.161), they are non-zero
only if k is a reciprocal lattice vector K differing from zero. This means that
the smoothness condition for U ( x ) may be expressed its

essentially non - zero for Ikl w ithin first 3 2


(3.35)
<< I(KlV)l, K # 0 , for Ikl outside f i r s t BZ
3.3. Effective mass equation 255

The left-hand xlde of the inequality (3.36) is of the order of magnitude of


the change of the perturbation potential U ( X ) over a primitive unit cell of
the direct lattice, that is, a1VUI with a as lattice constant. The right-hand
side of (3.35) can be estimated by a characteristic band structure energy, like
the fundamental energy gap E,q between the valence and conduction bands.
With this, the inequillity (3.35) takes the form

alVUl << E g . (3.36)


It states that, for a perturbation potential U ( x ) to be smooth and our first
condition to be fulfilled, the changes of U(x)over a primitive unit cell must
be small compared to the energy gap.
Secondly, we assume that the wavefunction $(x) can be set up exclusively
from Rloch functions of only one band YO, i.e., that $(x) should be of the
form

(3.37)

with F , ( k ) as a function which has yet to be determined. It turns out


that this requirement has no contradictions if the perturbation potential is
smooth.
The thzrd requirement refers to the function F,(k). It is assumed that
Fvo(k)differs from zero only for small k-vectors in the sense used in equation
(3.35).
Fourthly, it will be assumed that for small k-vectors, which according to
(3.35) have to be considered exclusively, the Bloch factors u W k ( x ) in the
Bloch functions p,k(x) can be approximately replaced by their values at
k=0:

U v o k ( X ) ZS .mo(x). (3.38)
In the following subsection 3.3.2 we will see that the two last requirements
can be justified when eigenstates of the perturbed crystal exist having energy
eigenvalues just above and below the edge of band vo, and if only these
eigenstates are considered. If, additionally, the band edge lies at k : 0,
the small energy deviations also correspond to small k-values, and relation
(3.38) holds approximately. The restriction of the location of the band edge
to k = 0 can be omitted, and the derivation procedure can also be applied to
band extrema other than the center of the first B Z , the only modification
being the replacement, of the R X center k = 0 by the non-central critical
point k,as well BY of k by k - k, in the corresponding equations.
The above four conditions will now be used to simplify the Schrodinger
equation (3.32). Applying relation (3.38) the wavefunction (3.37) may be
written as
256 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Tihe k-sum is the Fourier transform F,(x) of the function F,(k), whence
1
F,(x) = - CFvo(k)eik.x. (3.40)
& k
Therefore $ ( x ) may bc written as

$44 = ~ , O ( X ) ~ Y l ( X ) . (3.41)
The function F,(x) is termed the envelope functzon or, in short, the enve-
lope. The envelope function, by definition, i s a bmooth function. Equation
(3.41) means that the truc wavefunction + ( x ) is obtained by enveloping the
rapidly oscillating Bloch lactor uvgo(x) with the smooth cnvelope function
Fuo ( X ) .
The two ronditions conccrning the smoothness of U(x) and Fvo(x) make
it possible to represent the matrix element (uk(l/(vk)in the Schrodinger
equation (3.32) in a substantially simpler form. We rewrite this element
using thc Fourier representation (3.33) of U(x) and the product form of the
Bloch functions p v k ( x ) , obtaining

Furthermore, we transform the integral over the periodicity region into a


sum of integrals over the unit cells of this region. A lattice point R is
associated with each unit cell, so that the sum runs over all lattice points of
the periodicity region. Because of the lattice periodicity of the Bloch factors,
+
only terms of the form exp [i(k k - k) . R] remain under the lattice sum.
The summation over them can be executed easily. With K as an arbitrary
reciprocal lattice vector, it follows that

C .i(k+k-k)-R = G3 Sklfkl-k,K. (3.43)


R K
With this, expression (3.42) takes the form

(vklUlvk)
1
= - C ( k - k + KIU)Ci$(K), (3.44)
f i K
wherein the notation for the so-called Bloch integral

(3.45)
3.3. Effective mass equation 257

has been introduced (0 = G3Ro). The Bloch components of the wavefunc-


tion @(x) in the Schrodinger equation (3.32), (vkl$) and (vklg), differ
from zero only for small k and k because of the smoothness of the envelope.
Therefore we need the matrix elements (vklUlvk) only for small k and
k. With regard to the smoothness requirement for the potential U(x), the
terms in expression (3.44) for (vklUlvk) yield no significant contributions
to the Schrodinger equation (3.32) if K # 0, so that

1
(vklUlvk) N -(k - klU)C$t(O). (3.46)
a
Approximating the Bloch factors u,k(x) and U v l k t ( X ) by, respectively, ud(x)
and uU,o(x),and applying the orthogonality of the Bloch functions, the Bloch
integrals of equation (3.45) c : ~ : (follow
o ) as
c,,,
kk (0) = .6
t, (3.47)
Substituting this relation in expression (3.44) we obtain

(v kl U Ivk) w (kl U 1k)6,,, . (3.48)


With this relation, the decisive step in simplifying the Schrodinger equation
(3.32) for the perturbed crystal has been done - the coupling between the
different energy bands caused by the perturbation potential has been elim-
inated. The Schrodinger equation now decomposes into separate equations
for the various individual bands, and these equations can in fact be solved
by a one-band Ansatz of the form (3.37). The coupling between different
wavevectors remains in place. This can be processed relatively easily. We
employ all results achieved up to now in the Schrodinger equation (3.32),
obtaining the following relation for F,(k):

This equation will be transformed from k-space into coordinate space. To


this end we multiply by (1/a exp)( i k . x) = (xlk) and sum over k. The
first term on the left-hand side thereby becomes

C &,,(k)Fuo(k) (XIk) = Ev0 (-iVz) Fvo(x). (3.50)


k

One can easily prove that this transformation is correct by expanding E,(k)
in a power series and by using the identity
258 Chapter 3. Electronic structure of semiconductor crystals with perturbations

The second term on the left-hand side of equation (3.49) can be written in
the form

C ( x l k )(klU l kPm (k) U(x)Fm(x), (3.52)


kk
and the right-hand side of (3.49) inluiediately becomw E F , ( x ) . With this,
one finally gets the following equation for F,,(x)

If one further replaces Evu(-iVz), in accordance with equation (2.204), by


a quadratic function of -iVz7 which is justified because of the smoothness
of F , ( x ) , and assumes an isotropic effective mass m+,, then it follows that

h2
E,(-iV,) = E,(O) + -(-iVx)2,
2mg
(3.54)

and

h2
[ L . d O ) - -a2
2m&
1
I U ( x ) F,(x) = E F , ( x ) . (3.55)

With the derivation of this equation, initially conjectured and now verified,
it, is quite clear that the influence of the perturbation potential can be ap-
proximakly determined from an equation in which the periodic potential
no longer appears and the effective mass replaces the free electron mass.
Equation (3.55) is therefore the desired effective muss equation The eigen-
function Fvo(x) of this equation plays the role of an envelope function for
the Uloch factor uvoo(x)in equation (3.41) for the true wavefunction $J(x).
Equations (3.53) or (3.55) are also called envelope function equatioas in t,his
context .
The essential requirements involved in the derivation of the effective mass
equation were the smoothness of the perturbation potential, the smoothness
of the wavefunction and its composition of Bloch functions from only one
band, as well as the k-independence of the Bloch factors. These assumptions
are decisive, and are often not fulfilled. Short-range perturbation potentials,
for example, are not smooth. Nevertheless, the effective mass equation (3.55)
is a suitable instrument for the solution of a series of important problems
of solid state physics. Point perturbations with smooth potentials are only
one example of this. Other problems which can be solved with the help
of the effective mass equation include artificial superstructures in a crystal
and external electric fields. These will be treated later in section 3.7 and
3.8, respectively, Also, the Coulomb attraction between electrons and holes
- which results in the formation of excitons as pointed out in Chapter 2 -
3.3. Effective m a w equation 259

may be treated by means of this equation. To account for external magnetic


fields, the effective maw equation has to be modified in a way which will be
indicated later in section 3.9.

3.3.2 Multiband effective mass equation


The oneband effective mass equation in its general form (3.53), derived
in the preceding subsection, completely solves the eigenvalue problem for an
electron in a crystal in the presence of a smooth perturbation potential. The
practical use of this equation depmds, however, on the dispersion law E,(k)
of the band under consideration. In the vicinity of the minima or maxima
of non-degenerate bands of cubic crystals one has a purely parabolic and
isotropic k-dependence, and the effective mass equation is, as we have seen,
no more difficult to solve than an ordinary Schrodinger equation with an
external potential. This picture changes for bands which display degeneracy
in the extreme. As demonstrated in section 2.7 on k p-theory, one then
in general has non-parabolic and non-isotropic dispersion laws. This does
not involve a major difference if the one-band effective mash equation can be
solved in k-space, i.e. in the form of equation (3.49). IIowever, for various
reasons it can be necessary to transform the effective mass equation into co-
ordinate space and to solve it there. This applies if basis functions other than
plane waves are better adjusted to the symmetry of the perturbation poten-
tial as, for example, the Coulomb potential of an impurity atom, or if no
perturbation potential is present, but the perturbation is introduced through
boundary conditions, such as in the case of artificial superstructures like su-
perlattices and quantum wells. Because of the non-parabolic and anisotropic
dispersion laws in the case of degenerate bands, the effective mass equation
(3.53) in coordinate space is more complicated, in particular higher order
differentia1 operators occur which make solution of the eigenvalue problem
practically impossible.
A resolution of this situation is offered by k . p-perturbation theory. As
we have seen, the parabolic and anisotropic dispersion laws of degenerate
bands occur in this theory through diagonalization of the matrix of the
Hamiltonian with respect to a basis set of Bloch functions Ivk)', which are
exact up to the first order in the k . p-perturbation (see formula (2.339)).
The elements of this matrix are relatively simple linear and quadratic func-
tions of the components of k. In the case of degenerate bands this sug-
gests, therefore, not to take a onecomponent effective mass equation as
starting point, but a multi-component one which is obtained by writing the
Schrodinger equation for the perturbed crystal in the approximate Bloch
basis 1vk)l.
We will undertake this program now. Regarding the perturbation poten-
tial U ( x ) and the envelope function F ( x ) , we pose the same requirements
260 Chapter 3. Electronic structure of semiconductor crystals with perturbations

as in the previous subsection - both should be smooth in the sense that


they change much less over a primitive unit cell than the periodic crystal
potential V ( x ) does. For the wavefunction $(x) of the perturbed crystal,
the expansion with respect to the approximate Bloch functions reads

?clW= cuk
(.kl$)(xtvk), (3.56)

and the Schrodinger equation (3.29) in this representation takes the form

(vklH + Ulvk) (vkl$) = E(vk($). (3.57)


uk
The matrix (vklHIvk) of the unperturbed Hamiltonian H is diagonal
with respect to k, k, and block-diagonal with respect to the band indices
v , v with the blocks referring to degenerate bands. The diagonal elements
(vk(U(vk) of the perturbation potential U may be calculated by means
of relation (3.48) which gives the elements of U with respect to approximate
Bloch functions their Bloch factors u v k were replaced by u d . In the terminol-
ogy of section 2.7, these are Bloch functions in zero-order k . p-perturbation
theory or Luttinger-Kohn functions. Using relation (2.339), which expresses
the first order Bloch functions Ivk) in terms of Luttinger-Kohn functions, it
follows that the diagonal elements (vklUlvk) arc the Fourier transforms
(klUJk),just as in relation (3.48) of subsection 3.3.1. However, unlike this
ielation, thc matrix (vklU lvk) has also lion-vanishing off-diagonal ele-
ments with respect to I / , 11. This is due to the k-dependence of the Bloch fac-
tors in )vk)l which has been omitted in subsection 3.3.1. The non-vanishing
off-diagonal elerrienls (vklU)vk) may he calculated in the same manner
as the diagonal elements before. One obtains

The factor at (klVWlk) in equation (3.58) has the order of rnagiitudr of


the lattice constant a, as can easily be seen replacing p by x by means of
Heisenbergs quation of motion. This implies that the off-diagonal elements
of U ( x ) are of the order of magnitude of the relative change of U ( x ) over
a primitive unit cell. Terms of this order of magnitude are to be neglected
within the framework of effective mass thpory. Thus the matrix elements of
// with respect to first order Bloch functions Ivk) are approximately given
by the relation

(vkl U Idk) = (klU Ik)buu,, (3.59)


which corresponds to relation (3.48) of subsection 3.3.1. Note lhat equation
(3.59) is valid independent of whether the two bands v and Y arc dcgeneratc
3.3. Mective mass equation 26 1

or not; even degenerate bands are not coupled by perturbation potentials U


if these potentials are smooth.
Before simplifying the Schrodinger equation (3.57) still further for the
actually interesting case of degenerate bands, we will examine the case of
non-degenerate bands.
Non-degenerate bands
WTeassume that p(x) can be expressed in terms of the approximate Bloch
functions of only one band v g ? so that

'(vkl@)= b,,F,(k). (3.60)


With the Ansatz (3.60), vj(x)may be written as

$(x) = CFw(k)(xlvok)l. (3.61)


k
Employing expression (2.334) for the approximate Bloch functions jvok)', it
follows that

where &(x) is the Fourier transform of F,(k) in coordinate space. The


second term of @(x) of equation (3.62) does not occur in the corresponding
equation (3.41) for +(x). This term is again due t o the k-dependence of
the Bloch factor in Ivk)', which was neglected in subsection 3.3.1. In fact,
its relative magnitude with respect to the first term is of the order of the
relative change of the envelope function F , ( x ) over a unit cell. Again terms
of this order of magnitude are again to be neglected within the framework
of effective mass theory. We therefore obtain the same expression V(x) as
in equation (3.41) above.
The Hamiltonian matrix '(vklHlv'k')' in (3.57) is approximately diag-
onal with respect to the band indices, and the diagonal elements are the
eigenvalues Ez(k) of H calculated in section 2.7 in second order k. p per-
turbation theory. Employing expression (2.336) for E?(k),we obtain

where hi&8 are the elements of the reciprocal effective mass tensor accord-
ing to formula (2.337). Using the Ansatz (3.59) for '(vklp) and expression
(3.631 for l(vk)H]vk)', the Schrodinger equation (3.57) yields
262 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Fourier-transforming this equation into coordinate space, we have

This equation is in agreement with the effective mass equation (3.53) if the
latter has E,,,(--zV) replaced by a parabolic expression in the components of
-iV,refraining, however, from the assumption of isotropy of the band struc-
ture. If the effective mass tensor reduces to a scalar quantity ( l / m ~ & ) 6 ~ 0 ,
then (3.65)yields the effective mass equation in the form (3.55). The deriva-
tion of the effective mass equation based on k . p perturbation theory thus
produces this equation automatically in parabolic approximation. In the ear-
lier derivation of subsection 3.3.1, an effective mass equation was obtained
that did not yet contain this approximation, which was invoked only later.
The advantage of the derivation of the effective mass equation within the
framework of k * p-perturbation theory is clearly manifested when degener-
ate bands are considered, which we will address next.

Degenerate bands
We direct our attention to the valence band maximum of semiconductors
with diamond type structure (see section 2.7). Accordingly, we assume that
at k = 0 a degenerate band level E,(O) exists with either symmetry rhs
or r: depending on whether spin is omitted or not. First we consider the
case without spin. The three degenerate ,b'T Rloch states of energy E w ( 0 )
arc- distinguished, as before, by an integer index m. The Bloch functions
at k = 0 arc therefore written as ) o m O ) . If we are interested in eigcnstates
$(x) of the Schrodinger equation of the perturbed crystal whose energy
eigenvalues are expected to be near the valence band edge, we can represent
$(x) as a superposition of the Bloch functions Ivrnk)' in first order k . p
perturbation theory. Thus, we set

(3.66)

As above, in the case of non-degenerate bands, it approximately follows that

(3.67)
3.3. EEective mass equation 263

with F,(x) as the Fourier transform of F,(k) in coordinate representation.


Now, expression (3.67) for '(ukl$) is substituted into the Schrodinger equa-
t.ion (3.57). The subspace of valence band states is approximately decoupled
from the remainder of the Hilbert space. Using equation (2.341), the matrix
elements '(vklHlvm'k')' may be written as

where the coefficients D;:, are defined in equation (2.342). The matrix ele-
ments '(vklUlvrn'k')' of the perturbation potential U follow from equation
(3.59). They vanish if v # wm'. For v = urn' they differ only from zero if
m = m'. With this, the Schrodinger equation (3.57) takes the form

Transforming to coordinate space and taking account of relation (3.51), the


equation

+$
+ Ci(X)h,,!
1 Fm'(X) = E F m ( x ) .

@
(3.70)

($
follows. Using the notation D for the fourth-rank tensor D:fm, I
*
~ ~ l ~ =

for the unit matrix 6,i, and F (x)for the vector (F,(x),F,(x).F,(x)),t,his
equation may be written in matrix form as

(3.71)

In the case of the &I valence band the Hamiltonian matrix is given by rela-
t,ion (2.344). With t,his, the envelope function equation then reads, explicitly,
264 Chapter 3 . Electronic structure of semiconductor cry.staLs with perturbations

Here, derivatives of the type 62/6z2 are denoted by symbols like a;, and
those of type a 2 / d r d y by symbols like a&.
Consideration of spin leads to two changes. Firstly. one has to use spin-
dependent Bloch functions (sxlvmak0)' and, therefore, also spin-dependent
envelope functions Frrw(x, s ) , and secondly the spin-orbit interaction oper-
ator H,, is to be added to the Hamiltonian operator H . Thereby, Hk.p
U
becomes Hk.=, which has the consequence for D that the matrix elements of
p are replaced by those of ii from equation (2.353). If interference terms be-
tween the k ' p-interaction and the spin-orbit interaction H , are neglected,
the representation of the latter in the approximate Bloch basis leads to a
*
k-independent tensor Hs,.The envelope function equation then reads

It is advantageous to first transform this equation into the angular momen-


U
tum basis l j m g ) ,in which the operator fiso is diagonal, before one proceeds
to solve it. Using the arguments of section 2.7, and assuming a sufficiently
large spin-orbit splitting energy A, the coupling between the I'i and r$
valence bands may be neglected. For the T i' valence band, i.e. with j = $,
one arrives at the following $-component envelope function equation:

The quantities d,R , $, here are differential operators defined as follows:


1
Q = --(L
2
+ nf)(a: + a;) - Ma:, (3.75)

+ =
1
--(L+ 5n/I)(d2+ a;)
6
- -(2L
1
3
+Mp:, (3.76)

s* = i-(& N - ia&, (3.77)


v%
(L- M ) ( a2, - a,)2 - 2iNa,B,] . [3.78)

The energy origin in (3.74) was moved to the valence band maximum, which
was previously located at A/3. The envelope equation (3.74) makes it pos-
sible to calculate acceptor states and hole bands in superlattices of the dia-
mond and zincblende type semiconductors. One can derive a similar equation
within the Kane model of the rs-valence-band-rs-conduction band complex
of zincblende type semiconductors.
3.4. Shallow levels. Donor and acceptor states 265

3.4 Shallow levels. Donor and acceptor states


We will now explore stationary states of a semiconductor crystal caused
by point perturbations which givv rise to smooth long-range perturbation
potentials V'(x). The most important point perturbations of this kind are
isocoric substitutional impurity atoms, i.e. atoms from rows and columns of
the periodic table which are close to the row and column of the host atom
which is substituted. Below we will concentrate on such point perturbations.
We denote the difference of the core charge numbers of the impurity and host
atoms by

AZ = Z I - XI?, (3.79)
and assume that it is non-zero. The sign of AZ can be both positive as
well as negative. In this sense we can speak of positive and negative point
perturbations the point perturbation is positive if the core charge number
of the impurity atom is larger than that of the host atom, negative, if it
is smaller. For the pcrtiubation V,'(x) of the core potential, we m a y use
expression (3.11) which here we write as

(3.80)

In general, V,'(x) does not yet represent the entire perturbation potential in
the one-particle Schrodinger equation (3.21) for the perturbed crystal. One
still must add the Hartree potential Vh(x) and the exchange potential Vfr(x)
caused by other electrons localized at the center. For impurity atoms with
only one valence electron more or less than the host atom, these potentials,
as a rule, have the effect that only one electron or hole can be bound at the
center. For this one electron or hole, the Hartree and the exchange parts of
the perturbation potentials vanish, i.e. V,'(X) is itself the entire perturbation
potential, and the one-particle Schrodinger equation has the form (3.22). For
impurity atoms with lAZl > 1, Vb(x) and Vfr(x)do not vanish. The effect
of a non-vanishing VA(x) will be discussed later.
To solve the one-electron Schrodinger equation (3.22) with the potential
V,'(x) of (3.80), we may use the effective mass theory derived in section
3.3. Here, effective mass equations must be written down for those critical
points of energy bands where changes of the spectrum of energy eigenvalues
dae to the perturbation potential are to be expected. In Chapter 1, it was
pointed out that such changes occur at the bottom of the conduction band
and the top of the valence band - for impurity atoms with A 2 > 0 discrete
levels appear in the energy gap just underneath the conduction band edge,
266 Chapter 3. Electronic structure of semiconductor crystaIs with perturbations

and for impurity atoms with A 2 < 0 one has such levels in the energy
gap just above the valence band edge. Effective mass equations are needed,
therefore, both for the conduction band edge as well as for the valence band
edge. In writing down these equations we start with a simple twc-band
model. The minimum of the conduction band (vo = c ) , and the maximum
of the valence band (v, = w) are assumed to be non-degenerate and to lie
at k = 0. Furthermore, the k-dispersion of the two bands is assumed to be
isotropic and parabolic in the vicinity of k = 0, whence we have

(3.81)

Ti2
&(k) = -----k2, (3.82)
2m;
with rn; and ml, as effective masses of electrons and holes, respectively. In
reality, for many semiconductors such as Si and Ge. the conduction band
edge does not lie at k = 0. The valence band edge zs found at that point for
many semiconductors including the ones mentioned; but there is degeneracy
between heavy and light hole bands at k = 0. Nevertheless, we will initially
proceed with the above simplifying assumptions. The relevance and neces-
sary improvements of this idealized model are yet to be discussed separately.
For reasons which will become clear below, the model is often referred to as
the hydrogen model

3.4.1 Hydrogen model


For the hydrogen model we have the two effective mass equations

(3.83)

(3.84)

In the usual form of a Schrodinger equation they read

(3.86)
3.4. Shallow leveb. Donor and acceptor states 267

Apart from notation, signs, and constant factors these are the Schrodinger
equations for charged particleu in the potential of a point charge, just like
in the case of the hydrogen atom. From experience in quantum mechanics,
we know that h r an dttrtlclite potential, there is a continuum of positive
eiierpy eigmvalues and, in addition, there ale discrete energy levels which
occur for negative energies. Fur repulbive potentials there are only pusitixe
energy eigenvalues, but no bound statrs at negative energies. From this it
follows that for a positive point perturbation (upper sign in equations (3.851,
(3.86)), discrete e n e r a levels are to b e expected just below the conduction
band edge, and for negative point pertiirhations such levels are exppctpd
just above the valence band edge. Thew dibcrele levels lie in the emrgy gap
between the conduction and valence bands where, as we know, no energy
eigenvalues can appear in an ideal crystal. We consider first the cwe of the
conduction band.

Conduction band

Transferring results from the quantum mechanical treatment of the hydro-


gen atom to our present situation, we get the discrete energy levels & of
principal quantum niimher 71, n = I , 2 , . . ., as ( E , - E ~=) - ~ * / nor~

(3.87)

with
(3.88)

where Ef - (e'm//2ha) denotes the binding or Rydberg energy of the hydro-


gen atom. The m~av~fin~rtioii
FTdm(x) for principal quantum number n = 1,
angular mumedurn quantum number I 0 and magnetic quantum number
~

m - 0 reads

(3.89)

with
(3.90)

where a g = ( h 2 / e 2 m )denotes the Bohr radius of the hydrogen atom. The


orders of magnitude of the energy Eg and the length ag can be estimated
easily. With E g FY 13.6 eV,:a = 0.5 A , ( m z / m )= 0.2, c = 11.4 and =
1, it follows lhat

(3.91)
268 Chapter 3. Electronic structure of semiconductor crystals with perturbations

a g --" 0.5 (=)


11.4
AM 29 A. (3.92)

Compared with the width of the energy gap which is of order of magnitude
1 el/, the discrete energy level El lies closely below the conduction band
edge (see Figure 3.4). Even closer are the states with n = 2 , 3 , . . .. The des-
ignation shallow levels is used in this situation. The wavefunction for energy
level El decays exponentially with distance from the core of the impurity
atom with the characteristic decay length a g . In contrast to band states,
which are evenly spread out over the entire crystal, we have, therefore, lo-
calized electron states (see Figure 3.4). The localization radius a g is clearly
larger than the distance between two crystal atoms, which is approximately
2.5 A. This means that the localization of an electron at an attracting
impurity atom is weak if considered on the interatomic length scale of the
crystal. Exactly this is to be expected considering the smoothness require-
ment made at the outset, which states that only Fourier components of small
wavevectors k should contribute to the wavefunction. We can also examine
the validity of this requirement quantitatively. Omitting a k-independent
factor, the Fourier transform of Fcloo(x)is given by the expression

(4.93)

At the edges of the first B Z one has k FY (../a) and a g k M w ( a g / a ) FY 10.


Between the center of the first B Z and its edge, Fcloo(k) therefore falls off
by about 4 powers of ten. The smoothness requirement is therefore fulfilled
very well. In Figure 3.4 the localization of the wavefunction in k-space is
also indicated.
Recalling the theory of the hydrogen atom, it is known that the effect of
an attractive Coulomb potential is not restricted to the formation of bound
states with energy levels in the previously forbidden negative energy region.
Changes also occur in the continuum of positive energy eigenvalues which are
allowed even without Coulomb potential. They are manifested in the fact
that the eigenstates, which were previously spherical waves, are scattered
by the Coulomb potential. Their amplitude at the positive center becomes
larger, and further away from the center it is smaller. This leads to a change
of the density of states in the continuous part of the energy spectrum (the
definition of the density of states is given in section 2.5). According to
Levinson's theorem, which we will discuss in the next section in somewhat
greater detail, this change takes place such that the total number of all states
in the presence of the perturbating potential, including the bound states at
negative energies, remains the same as the total number of states without
the perturbating potential. For each bound state of negative energy a state
of positive energy is therefore excluded from the energy spectrum.
3.4. S h d a w lev& Donor and acceptor fitates 269

AZ a 0 dZ.0

L1 O k r\ O k

Figure 3.4: Illmtration of shallow localized States in coordina1.e spat:e and in k-


space far mibslitutional impurity atoms with either posit.ive or negative differences
A 2 beOween the impuril,y core charge nunthers and the h o d core charge numhers
(upper part of figure). The occupation of the shallow levels by eI&rons is presented
in the lower part of the figure. The levels act as donors of elwtctruns for A 2 > 0 and
ss acceptors of electrons for A 2 < 0.

The abovr-Inmtionfd result,^ for shallow impurity levels appended to an


isotropic parabolic conduction band can be immediately iransferrd to an
isotropic parabolic valence band. Below, we will writs them down without
further derival ion.

Valence band
For a negative point perturbation there appear discrete energy lpvels in the
energy gap closely above the valence band d g r . The energies ol thew levels
are given by the exprcssiorrs

EB
En - -
z, ' (3.94)

(3.95)
The pcrtinent wavefunctions are localized at the perturbation center. The
wavefunction of the ground state, n = 1 and 1 = m = 0, has the form (3.89)
270 Chapter 3. Electronic structure of semiconductor crystals with perturbations

with

(3.96)

as the effective Bohr radius. In Figure 3.4 (lower part) the energy levels and
localization regions are shown schematically.
Although the population of energy levels by electrons will not be treated
systematically until the next chapter, we wish to deal now with a special
case, namely with the occupation of the shallow impurity levels discussed
above at absolute zero temperaturc, 7 = 0 K . This consideration will lead
us to a better understanding of the nature of these impurity states. First of
all, wc assume the case of a positive point perturbation, i.e., the core charge
number 21 of the impurity atom is taken to be larger than the core charge
number Z H of the host atom. For simplicity, we assume 21= ZH 1. Then+
the impurity atom also has one valence electron more than the host atom.
To be specific, we can imagine it as a P atom in a Si-crystal. According to
Levinsons theorem, which mandates that the total number of states must
remain the same both with and without perturbation, one can proceed on the
assumption that nothing will change due to the impurity atom regarding the
number of states in the valence band. Therefore, all the valence electrons of
the host atoms and all those of the impurity atom, except for the additional
electron, can be placed in the valence band. The additional impurity electron
has to reside in the lowest unoccupied energy level. That is the shallow n =
1-level just below the conduction band edge which, according to the above
results, arises from the impurity atom. The resulting occupation is shown
in Figure 3.4. If the temperature is increased this electron can easily be
excited from the shallow level to the conduction band. There, it is no longer
localized but spread out uniformly over the whole crystal. representing a
freely mobile negative charge carrier. The shallow energy level therefore
functions as a donor of a free carrier. One calls it a donor level and the
impurity atom itself as a donor. One can also say that a donor atom (with
AZ = 1) has one surplus valence electron which is bound relatively weakly
and can be transferred easily by thermal excitation to the conduction band.
The impurity atom remains in the single positively charged state, i.e., it is
singly ionized. For A 2 2 2 one has doubly or multiply ionizable donors. We
will discuss this later more exactly.
Next, we consider an impurity atom with Z I = Z H - I, i.e. with one
positive elementary charge in the core and, with this, also one less valence
electron than in the host atom. Due to the substitution of a host atom by
such an impurity atom, the number of states in the valence band decreases
- according to Levinsons theorem by one if only the n = 1-level is counted.

With this, the number of valence band states is still large enough to host all
3.4. Shallow Jevels. Donor and acceptor states 271

valence electrons, but, also, there remain no unoccupied states in the valence
band. The n = 1-level in the energy gap remains empty at T = 0 X. If the
temperature is increased, an electron from the valence band can easily be
excited into the impurity level. The level accepts an electron, it functions
as an acceptor. In the valence band itself, a hole remains which, as we
know from Chapter 1, behaves like a positive freely mobile charge carrier.
Consequently, we can also say that, under thermal excitation, the acceptor
transfers a hole to the valence band, whereas at T = 0 K the hole was
bound to the acceptor. This picture corresponds to the model of a negatively
charged point perturbation which binds a positive hole and transfers it to
the valence band under excitation. The analogy to the positively charged
donor center which binds an electron and donates it to the conduction band
is obvious. Similar interpretations are available for doubly and multiply
ionizable acceptors.

3.4.2 Improvements upon the hydrogen model


As already indicated, the above hydrogen model of shallow impurities, for
which the band extrema lie at the I? point and the bands are non-degenerate
at r as well as isotropic and parabolic in its neighborhood, cannot be directly
applied to many semiconductor materials including Si. Moreover, as we
know from section 3.2, the perturbation potential V(x) has, in general. a
short-range part as well as an electron-eIectron interaction part which are
not considered in the hydrogen model. Below, we treat corrections to the
hydrogen model which can be traced back to the cited effects. Some of these
corrections appear only at donors. others only at acceptors. Therefore, we
treat the two kinds of impurities separately.

Donors

The conduction band of Si and other indirect semiconductor differs in three


ways from the simple isotropic hand model. First, the minimum lies out-
side of the center of the first BZ;second, for symmetry reasons, one min-
imum carries the implication of several equivalent minima or valleys; and
third, the band structure in the neighborhood of each individual minimum
is anisotropic. While the off-center location of band minima has no direct ef-
fect on the donor binding enerm, the many-valley feature and the anisotropy
of the band striiccture rln have such an effect. Due to the existence of several
valleys (dist.inguished by a valley index i ) , the wavefunction 01 the donor
state $(XI has to be expanded with respect. to the appr0ximat.c Bloch func-
+
tions (xlclri [k- kill1 for all mutu& degenerate minima ki, rather than
+
with respect to the approximate Rloch function (xlck, [k- kc])of one
minimum k, only Because of this, the one-component envelope function
272 Chapter 3. Electronic structure of semiconductor crystah with perturbations

F,(k) becomes a multi-component envelope function Fi(k), and we have

The envelope function equation in k-space reads

rx
k j
h2
-
2
[
(k - kilMClk - ki) 6ij6kkl + (klVlk)] Fj(k - kj) =
= EFi(k - ki), (3.98)
where Mi is the effective mass tensor of valley i.
The wavevectors k - ki and k - k j vary in a small neighborhood about
the zero point, i.e. k and k are vectors near the respective minima ki and
kj. For i # j , the matrix elements (kJVJk)of the perturbation potential
V(x) can approximately be replaced by the k- and k-independent expres-
sions (kilVlkj). The latter can easily be transformed into coordinate space
where they result in an additional 6-function-like potential. With i = j , the
matrix elements (kJVJk)result in the perturbation potential V(x) which
is already present in the hydrogen model. The effective mass equation (3.98)
in coordinate space thus reads

= EFi(X). (3.99)
If V(x) is the screened Coulomb potential of equation (3.80), as was as-
sumed, the absolute values of the inter-valley Fourier components (kilV1kj)
are negligibly small in comparison with the intra-valley Fourier components,
because of the large wavevectors ki - k j occurring in the former combined
with the (l/lkl)-dependence of (klV) on k. However, as we know from sec-
tion 3.2, the true perturbation potential V(x) differs from the pure Coulomb
potential, because the screening of a point charge by the semiconductor is
wavevector dependent, and on the other hand, because V(x) contains a
short-range part, which cannot be traced back to the potential of a point
charge in any way. Both modifications can cause the inter-valley matrix
elements (hlVJkj) of V(x) to take larger values. Nevertheless, they still
remain so small that they can be considered by means of perturbation the-
ory. In zero-order approximation they may be omitted completely. Thereby,
one obtains a separate effective mass equation for each valley from (3.99).
In case of k, = ( O , O , k z ) , it reads, in coordinate space
3.4. Shallow levels. Donor and acceptor states 273

Analogous equations hold for the other valleys. This means that, neglecting
inter-valley coupling, all valleys give rise to the same donor level.

Anisotropic effective masties

In contrast to the hydrogen model, the effective mass of the envelope q u a -


tion (3.100) depends on direction in k-space. In this circumstance, analytical
solutions arc not available, and one mist resort to approximations and nu-
merical calculations. One possible approach is the application of a variational
procedure, in which the eigenfunctions are represented as linear combinations
of niembers of a set of auxiliary functions whose general shapes are adjusted
as well as possible to the real solution, but which still contain free parame-
ters. Using this representation one calculates the energy expectation value
of the Ilamiltonian and varies the parameters until its absolute minimum is
reached. This minimiim value then yields an approximation for the ground
state energy, which is better as the linear combination of auxiliary functions
approximates the actual eigcnfunction. The success of this approach thus
depends in an essential way on the auxiliary functions used. They should,
in particular) correctly reflect the symmetry of the actual eigenfunctions, in
the case considered here the symmetry of the ground state wavefunction.
One can also employ the variational process for the calculation of excited
states. In this case the wavefunction of the first excited state must, in ad-
dition, be orthogonal to the ground state wavefunction, that of the second
excited state must be orthogonal to both previously calculated states, etc.
Numerical results obtained by means of this procedure for the energy levels
of the P-donor in Si are reproduced in Figure 3.5. For a simple estimate of
the anisotropy effect, one can replace the reciprocal effective mass rn: of the
hydrogen model by the reciprocal effective mass (l/m:ll + 2 / m ; ) / 3 obtained
by averaging the direction-dependent reciprocal effective masses of (3.100)
over all directions in k-space. With this, one obtains energy levels for the
simply ionizable donor in Si, which are shown in Figure 3.5a. The binding
energy E B is represented as 29 meV in this approximation.

Inter-valley coupling

As a second corrective step, the inter-valley coupling in the effective mass


equation (3.99) will be considered. In perturbation theory, the calcula-
tion of eigenvalue corrections involves the diagonalization of the matrix
O(kilVlkj)lFj(0)12. Using the example of Si, this calculation will now be
274 Chapter 3. Electronic: structure of semiconductor crystals with perturbations

Figure 3.5: Ground state and


first excited states of a phos-
phorus donor in Si: a) hydro-
gen model; b) including effcc- ->T 0'0
tive mass anisotropy; c) includ- 2! -0,Ol
ing central cell corrections, effer- CI,
w
live mass anisotropy arid inter- I
w
valley coupling; d) exxperimen-
tal results. ( A f t w Rlakemore, -0,02
1985.)

- 0.03

-0,04

-0.05

performed explicitly._ In this


- case~ one has 6 minima, which lie on the cubic
axes k,, k,, k , and k,, k,, k , close to the respective X-points. From symme-
try considerations it follows that fl(k,lV'lk,)II;k(0)12 has the general form

O b b c b b
b O b b c b
b b 0 b b c
c b b O b b 7 (3.101)
b c b b O b
b b c b b o

b = fi(k~lv'lky)l~z(0)12,
r = ~t(kzIv'lkz)l.)IFz(0)12 (3.102)
_ _ Lines and columns in (3.101) are written in the sequence
have been-used.
k,, k,, kzrk,, k,, k,. The matrix (3.101) has the three eigenvalues

A&I = 4b + C, 1 - fold
AEz = -2b + C, 2 fold
AEg - -c, 3 - fold (3.103)
3.4. Shallow lwels. Donor and acceptor states 275

with the degrees of degeneracy indicated. Thus, inter-valley coupling par-


tially removes the 6-fold degeneracy among the valleys.
The sphtting of the ground state donor lrvel in Si due to inter-valley
coupling can also be predicted j u s t oti the basis of group theory. One con-
siders the representation of the symmetry group of the envelope e q u e h n
(3.99) for si, i.e. __ o h ) in the space of the 6 reciprocal vector compo-
of _
nents k,, kg,k,, k,, k,,, k,. This representation is reducible. To demonstrate
this one may employ the transformation rules of the vector components
x , y, z , a,7j, Z under the action of the clcments of o h (summarized in Table
A.7 of Apprndix A), realizing that these component? transform in the same
way as thc reciprocal vector components considered here. Then one easily
finds that the irreducible parts of this representation arc rl ( A , ) . r12 ( E )
and Iz5(Tz). The notations A , E , Tz are commonly used in the theor,- of
l o c a l i d states, and WP will also employ thcm here. With this. it is clear
that the symmetry of the AEI-state is A l , that of the two AETstatcs is E ,
and that of the three AE3 states is 12.
The size of the splitting depends on the inter-valley nitrtrix elemen1 of
the perturbation yolmtial. If one considers only thc first of thc two above-
mentioned effects which result in matrix elernentv ol considerable size. i.e.
the wavevector dependence of screening, and assumes that screening has
become fully ineffective at large wavevectors k, - kj,therefore replaring the
screened Coulomb potential V(x) of equation (3.80) by the bare Coulomb
potential -(e2/IxI) in (ktlVlkj), then it follows that

(3.104)

With this, one obtains b as -2.36 mel and c as -1.2 met. This yields
AEl = -10.8 meV. A E z = 3.6 met and AE3 = 1.2 meV.
A 3-fold splitting of the P-donor ground state level in Si has, in fact,
been observed experimentally (Aggrawal and Ramdas, 1965). The simple
theoretical estimate presented above agrees remarkably well with the ex-
perimental splittings of 11.6 meV between the A l - and T2-levels, and of
1.3 meV between the E - and T2-levels (not resolved in Figure 3.5). The
results of a numerical treatment of inter-valley coupling shown in Figure 3.5
are even closer to the experimental values. However, the agreement is not
as good for the absolute position of the ground state level. The addition of
A E l , i.e. of 10.8 meV to the binding energy of 29 meV without inter-valley
interaction yields a corrected binding energy of 39.8 meL which is closer to
the experimental value of 45 meV, for Si:P but still clearly below.

Chemical shifts

In Table 3.5 we list experimental binding energies for a number of singly


276 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Table 3.5: Experimental binding energies of several shallow donors and acceptors
in Si, Ge and GaAs. (After Landoldt-Bornstein, 1982.)

Material

Si
Sb Ga
In 153
P 13 B 11
Ge As 14 A1 11
Sb 10 Ga 11
Bi 13 In 12

GaAs

ionizable donor and acceptor atoms in Si, Ge and GaAs. Not only are the
absolute values of these energies striking, but also the fact that they are
not equal for donor or acceptor atoms of different chemicals stands out.
Contrary to the prediction of the above theory, there are dependencies of
binding energies on the chemical nature of the impurity atoms. One refers
to these as chemical shifts.
The absence of chemical shifts in the calculated binding energies is a
consequence of the approximation of the perturbation potential V(x) as a
purely Coulombic potential in the hydrogen model. As we know from section
3.2, the correct perturbation potential of an impurity atom also contains a
short-range part, to which all central cell corrections contribute. Among
these, there are also contributions which depend on the chemical nature of
the donor. The latter are the main reason for the experimentally measured
chemical shifts of the donor binding energies. The effects of central ceE
corrections are particularly large in the case of the 1s-ground state, as one
can recognize from Figure 3.5 for the special case of the P-donor in Si. A
weaker influence, and, accordingly, better agreement between experiment
and the predictions of a theory omitting central cell effects occurs for the
excited states with n = 2 , 3 , . . .. This is to be expected since the excited state
wavefunctions have maxima that do not lie at the perturbation center x = 0
3.4. ShalIow levels. Donor and acceptor states 277

(which does happen in the case of the ground state) but further outside.
Electrons in the excited states are therefore less affected by changes of the
potential in the central cell than are electrons in the ground state.
Generally, the occurrence of pronounced chemical shifts means that the
perturbation center is no longer shallow and the effective mass theory is no
longer applicable for its theoretical treatment (see section 3.5 for further
discussion).

Acceptors

Corrections to the simple hydrogen model are also necessary for acceptors.
Of course, the maximum of the valence band lies in the center r of the first
B Z for all diamond and zincblende type semiconductors, as is assumed in
the hydrogen model. However. this maximum is degenerate: depending on
whether spin-orbit interaction is important or not one has, respectively. the
%fold degeneracies of the representations l'b5 (diamond) or (zincblende)
in the space of scalar functions. or the 4-fold degeneracies of the representa-
tions I?$ (diamond) or Ts (zincblende) in the space of twc-component spinor
functions. In the vicinity of r, this degeneracy splits off and one has ihree
or two anisotropic valence bands. In section 2.7 these were approximated by
two isotropic parabolic bands. one for heavy holes and one for light holes.
If one also uses this approximation here, then there are two hydrogen-like
series of acceptor le\-els, one for each sort of holes. Calculating acceptor bind-
ing energies by means of expression (3.95) in the case of Si yields 50 met7
for heavy holes which are expected to form the ground state. In the case
of Ge. the result is 17 meV for heavy holes. One can hardy expect these
values to be in agreement with experiment. In fact, they differ apprecia-
bly from the measured ground state binding energies shown in Table 3.5.
Evidently. non-parabolicity and anisotropy of the band structure play an
essential role and must be taken into account in realistic calculations. This
may be done by means of the multfband effective mass theory developed in
the preceding section. Below. we will explain the application of this theory
to simply ionizable acceptors in Si. Owing t o the small spin-orbit splitting
energy A of 44 meV and the relatively large acceptor binding energies E B
(68 me\' for the isocoric impurity atom Al). spin-orbit interaction initially
will be neglected. Later this approximation needs to be corrected since E g
is not much larger than A. Without spin and spin-orbit interaction the
wavefunction ~ A ( x ) is a linear combination of the three Bloch functions
(xlumO),m = x. y, z of the representation Ti,, which are enveloped by the
*
components F,(x) of the envelope function vector F (x). Thus

(3.105)
278 Chapter 3. Electronic structure of semiconductor crystals with perturbations

The corresponding effective mass equation of the Ib5-valence band is given


by relation (3.72). If, therein. U ( x ) is identified as the Coulomb potential.
(3.80), one obtains, in the case A 2 = -1,

The eiigenvalues and eigeuvectors of this complicated set of equations cannot


be obtained, of course, in closed analytical form. One has to use spproxinia
tions and numerical calculations. The variational method again represents
a reasonable way to proceed. In this approach one starts with a choice of
suitable auxiliary functions for the expansion of the multi-component enve-
=$
lope fundion F (x). This function should have symmetry properties which
fit the symmetry of the acceptor state @ A ( X ) to be calculated. The mean-
ing of this will be explained below. We consider a particular acceptor level
E n , which in general will be degenerate. It belongs to an irrdiicible reprc
sentation r A of the point moup oh. The corresponding aigenhnctions are
. . ., and the pertinent multi-component envelope
denoted by $ A ~ ( x ) ,@m(x),
* *
function by ~ , . q l(x), FA? ( x ) ,. W e have to find the representation D A of
=$-
O h to which the set of vectors F A ~ (x),F A ~. ., . must belong in order that
the wavefunctions T , ~ A ~ ( x$~z(x):.
), , . formed from these vectors by meam

of equation (3.105), actually transform according to the representation r ~ .


3 *
The FA^ (x). FA^ (x),. . . belong to a space which differs from the ordinary
Hilbert syaccr in that it is riot spanned hy oue-comyonent basis funrlious, but
by three-component basis functions. A representation in this space is given
by the product of two represt?nt,atinns, one of which corresponds to the 3D
rqresentation according to which the components of each of the three-
= $ *
component, fiinctions FA^, FA^ transform separately, and the other is the
representation according to which the threecomponent functions transform
among each other. One can easily show that the latter representation must
be r A , so that $,.ql(x),+,.q2(x), . . ., forincd according to relation (3.105),b e
long to the representation r A , as has been supposed. For the representation
* *
D A of FA^ (x):FA^ (x),. . . it follows that

(3.107)
3.4. Shallow levels. Donor and acceptor states 279

This representation is reducible. The irreducible components determine the


symmetries of the three-component basis functions to be used for the r e p
resentation of the threecomponent envelope function of the acceptor state
under consideration.
Equation (3.107) yields a remarkable conclusion concerning the symme-
try of acceptor states. An envelope function basis vector which transforms
according to the unity-representation occurs in the expansion of the accep-
tor eigenfunctions $ A ~ [ X ) , @ ~ ( X .) .~only
. when the symmetry r-4 of these
functions is I'h5. The basis vector belonging to the unity-representation is
the only one which does not vanish at the central site x = 0. Since the wave
function of the ground state will also be non-zero at x = 0: D A must contain
the unity-representation if A is the ground state. According to (3.107) this
is only possible if the ground state has the symmetry Fh5. Acceptor levels of
other symmetry are necessarily excited states. The symmetries of all t h r e e
component basis functions involved in the construction of the ground state
envelope function are given by the relation

(3.108)
shown in Table A.27 of Appendix -4. Relation (3.108) determines the symme-
tries of the three-component auxiliary functions under rotations, therefore
the angular dependencies of these functions. In order to also get their ra-
diaI dependencies one expands the rl-, r12-, and ri5-functions with
respect to angular momentum eigenfunctions with the various quantum num-
bers 6, rn and multiplies the expansion coefficients of different values of I by
corresponding radial wavefunctions r1exp(-T/q). These are formed in anal-
ogy to the eigenfunctions for the Coulomb potential, but with &dependent
localization radii rl. The latter are treated as variational parameters, just
like the Coefficients of the auxiliary functions belonging to different irre-
ducible representations. Applying this procedure to the acceptor ground
state of Si, a binding energy of 31 me&' is obtained (Schechter, 1962). More
recent calculations, taking spin-orbit interaction into account, have resulted
in a value of 44 mel/ (Baldereschi and Lipari. 1977), which is very close to
the experimental value of 45 meV for 3 in Table 3.5 but, of course, does not
account for the pronounced chemical shifts seen in this table.
Calculations of acceptor binding energies have also been performed for
materials like Ge whose spin-orbit splitting energies are large compared to
the acceptor binding energies. Spin has to be taken into account under
such circumstances, i.e. the T'& valence band has to be replaced by the two
rg- and r$-bands. Owing to the large spin-orbit-splitting energy, however,
the spin-orbit-split I'F-band can be omitted. For the expansion of the 4-
component envelope function of the remaining rg-band one needs auxiliary
functions of + +
x I'i = ri fk + +
2r15 2r25 symmetry (see Table
280 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(A.28)). In the calculations for Ge, an acceptor binding energy of 10 m e v


is obtained, close to the experimental values of 12 nieV for In and 11 meV
for all other elements listed in Table 3.5. Obviously, chemical shifts are
very small in the case of Ge, unlikr Si where these shifts were found to
be very pronounced. The different behavior of acceptors in Ge and Si is
undmstandable if one looks at the absolute magnitudes of their binding
energies EB: in Ge, E D is siihstantially srrialler than in Si, reflecting the
fact that the effective heavy hole mass in Ge is smaller than in Si. This
implies that the acceptor wavefunctions of Ge have smaller amplitudes in the
central cell than those of Si. Thus the point charge Coulomb potential which
neglects central cell corrections giving rise to chemical shifts is expected to
work much better for Ge than for Si, as it dow in fact. For the same
reason, excited acceptor states in both Ge and Si are well described by the
multi-band effective mass theory using a point charge perturbation potential
(Balsderesrhi and Lipari, 1978).

Multiply ionizable donors and acceptors

Additional changes arc to be expected for substitutional impurity atoms


whose core charge numbers do not differ by & A 2 - 1 from those of the host
atom, as we assumed above, but differ by 2, 3 or more. Then 2, 3 or more
electrons or holes are weakly bound to the impurity atom at T - 0 K , and
1, 2, 3 or more electrons can be thermally excited into the conduction or va-
lence bands, leaving behind a 1-, 2-, 3- or morefold ionized impurity atom.
The interaction between the carriers bound at the center is described by the
Hubbard energy of equation (3.25). Without this interaction the effective
mass equat,ion for a non-degenerate isotropic band yields a hydrogenic s e
ries of energy levels with binding energy increased by the factor lA21' in
comparison with the corrcsponding energy for a simply ionizable donor or
acceptor. The Hubbard correction (3.25) results in a shiit of the energy lev-
els towards higher energies by an amount which depends on the number 'IL
of electrons or holes bound at the center.
We illustrate these remarks using thc example of the S-donor in Si. The
binding e n e r a without H a r t r e potential is four times larger than that of the
P-donor in this material. In the ground state, 2 electrons are bound at the
S-donor, both of which can be placed in the 1.9-level because of 2-fold spin-
degeneracy. If the Hubbard correction, which in this case is approximately
given by

is taken into account, then the 1s-level is shifted up by just this energy
Us.By exciting one electron into the conduction band, the neutral S-donor
3.8. Deep lewls 28 1

S(D+) bccarnes a single positively charged S(1+) donor. For the Is-lad
01 h e S ( l 1 ) ioii thr shift in energy by Us do= not occur. Thus this level
i s shifted down by I/, in comparison with the 1s-level of tbc S(0 +-) atom.
Exprrirneniully, one finds t,hat the ionk-ation energy of the neutral S(O+)
donor is 0.31 eV, and the ionization energy of the S(1t) donor j s 0.69 eV
(see Figure 3.6). From thefie values a Hubbard energy I i , of 0.28 eV may
be deduced. IIowever, both ionization energies are suhstantially larger than
the result. 4 x 0.03 el7 = 0.12 eb which follows from the hydrogen model.
Evident.ly, the central cell short-range poteniid cont,ribution is ewedial in
the case of the S-donor in Si. l l i s donor is a deep center rather than a
shallow me.

3.5 Deep levels

3+5.1 General characterization of deep levels


As we know from previous sc3ctions, hhc potential of a point perturbat.ion,
generally, consists of a long-range (Coulornbic] part and a short-rangp part..
In regard to their ability to bind states, the two potential contributions differ
substantially. In the case of the long-range Conlomb potential, hound $Pates
exist, for all possible potential strengths, i.e. for arbitrary magnitudes of the
point charge and dielect,rir:cobstant. Short-range potentials, on t,hecontrary,
must have a minimum st.rengt.h to he able to bind a stale. For example,
for a 3-dimensional potential box of depth Vo and radius a, the condition
Vo > h2/(8w>.uli) must he fulfilled for a b a u d state to exist (see, e g . ? Schiff,
1968). In this, Heisenbergs uncertainty principle manifests itself, in that
localization due to the potential leads to a non-vanishing expectation value of
momentum and 1,hus also of kinelic energy of the particle. Only if the dept8h
of the potential box exceeds the expectation value of this kinetic energy! can
t.hc potential prevent the partick from escaping the center. (Not.e, however,
that in the case of a 1-dimensional potential box, bound states exist for
arbitrarily small m d l depths sincc the 1ocalizat.ionand: with this, the average
kinetic energy decreases sufficiently- fast with decreasing well dept.h. This
indicates already at t,he outset that 1-dimensional models of deep centers
are ralher poor).
According to wction 3.4, shallow levek ocrur for statps which are bound
by a Coulomb potential, i.e., by the long-range contribution of the total
perturbation potential. The concornit,ant short-range potential part does
not suffice in this case for binding, it only leads to corrections of the binding
energy and the eigenfunction of the ground state, k1loa.n from section 3.4
as c e n l m l tell correcfioru. The spatial extension of the eigenfunction is
essentially given by the Bohr radius. If t.he strength of the short-range
282 Chapter 3. Electrunin structiire of Bemiconductor crystals w i t h perturbations

part of the potential increases, then a point will be reached at which the
short range potential is itself able to bind states. Provided the pertinent
binding energy i s Iargcr than that due l o the Coulomb potential, binding to
the center will be dominated by the short-range potential contribution, and
the chararter of the bound state will change. The spatial cxtension of the
eigenfunctjon of the ground state is then no longer dekrrninml by the Bohr
radius, but by thr lattice constant. The Coulomb potential only leads to
corrections to t h r binding energy and the cigmfunctions, which are mainly
determined by the short-range potential alone.
States which are primarily bound by the short range potential part, are
termed d w p . h e rorresponding la-& arc called deep levels, and the point
perturbations at which they occur, are called as deep centers. Literally, the
term deep level refers to an energy eigenvalue that lies deep in the energy
gap, far away from the two band edges, in contrast to shallow levels which
lie close to one of these edges. Actually. the location of deep levels deep
in thr energy gap is only a particularly important special case. as there
are also yet other case. - derp levels can also be close t o or even wzthm
one of the bands. The essential features of deep levels are their binding
by a short-range potential and, in combination with this, their localization
radius bring restricted to magnitude of the lattice constant. Because of their
strong localization, deep states resolve the spatially periodic fluctuations of
the rrystal potential while the eigenfunctions of shallow levels average them
out. The shallow levels can therefore be treated by means of an effective mass
equation. in the case of deep levels the requirements of effective mass theory,
namely smoothness of the perturbation potmtial and of the corrpsponding
wavefunction, are not fulfilled. Contrary to the assumptions of PEective mass
theory, these states cannot be s p t h e s i z d by Bbch functions of k-vectors
drawn from a small vicinity of a rritical point, and also not from Bbch
functions of only one-band. In a theory of deep states, therefore, neither the
effective mass is meaningful, nor can such a theory be based on a oneband
equation, not even a degenerate multi-band equation. Attempts t o retain
the effective mass concept and to consider solely ihe multi-band character
of deep level theory met with little success. In this context one can say that
skidow levels are eigenvalues of point perturbations which are capable of
description by means of effective mass theory (the concomitant short-range
potential contribution can be treated by rncans of perturbation thcorg in this
rase), while deep levels are eigcnvaliies whose treatment by effective mass
theory is impossible.
The discussion above is amply intuitive. it anticipates, in part, results
which are i n d d plausible, but which have yet to be proven rigorously.
This applies, in particular, to the fundamental feature which distinguishes
between shallow and deep levels, namely, that short-range potentials can
dominate binding only if they exceed a minimum threahold potential strength
283

2 1 0 -2 -4 -6

.3

Acceptors 2
.4 - - - _ aF:
.-

.5

.6 (eV)

2 1 -2 -4 -G -8 -1OleV)
Atomic Electronegativity

Figure 3.6: Ionization energy of substitutional impurity atoms in varioum tedra-


hedrally coordinate host semironductors as B function of the strength of the pcr-
turbation potential, measured in terms of valence shell s-level differences between
impurity and host atoms in the case of donors, and valence shell p-level differences
in the caae of acceptors. Donor energies are reIative to the conduction band edge,
and acceptor energies to the valence band edge. (After Vogl, 1981. Reproduced
from Boer, 1990.

- in the discussion above, the periodic potential of the crystal was omitted
from consideration. An experimental proof of the threshdd behavior of
short-range potentials in crystals may be taken from Figure 3.6. There: the
ionization emrgies of a series of substitutional impurity a t o m are plotted
as a function of the difference! between the atomic valence shdl energy levels
of the host and impurity atoms. This difkrence can serve as a measure of
the strength of the short-range potential. The fact that for donor states
284 Chapf,er3. Electronic structure of semiconductor crystals with perturbations

the diffrrence of 8-levels is taken, and for acccptor states the differenre of
p-levels, is not important foi the present discussion (we will return to this
point later). With increasing horizontal scparation of the impurity atom
from the host atom in the periodic table, the potential strength increases
more or less continuously. The ionization energies initially retain their small
valucs characteristic of shallow lcvels, Starting at a threshold distanre, they
suddenly become substantially larger, a manifestation of the fact that the
short-range potential has takm over binding and the level has become deep,
.Just like shallow levels, decp levels can also act as donors or acceptors
(sometimes even as both of them). Due to their mostly larger separation
from the hand edges they are, however, generally less effective than shallow
levels in enhancing the concentrations of free charge carriers. They exhibit
their greatest efiectiveness in just thp opposite process, the lowering of free
carrier concentrations for which, in turn, shallow levels are not very effective.
If, in addition to the neutral center, the single negatively charged center also
forms a deep level in the gap with sufficient separation from the conduction
band edge, the neutral center will capture electrons from the conduction
land, which are available there eithrr as equilibrium charge carriers due to
n-doping, or ~ L non-eyujlibiium
Y cariiers due to optical or other excitation of
the semiconductor. In the first case, one has a c o m p e ~ w u t i o nof the donors
by the deep center (for more 5ee Chapter 4), and in the second case, a cup-
ture of non-equilibrium electrons by the center (see Chapter 5). A center
which forms deep levels in the gap both in the neutral and simply posi-
tively charged state, plays an analogous role in regard to the compensation
of acceptors and the capture of non-equilibrium holes. Centers which can
capture both electrons as well as holes act as catalysts for the radiation-
less recombination of electron-hole pairs (see Chapter 5). If one wants to
enhance radiative recombination as strongly as possible, such as in semi-
conductor light emitting diodes or laser diodes, non-radiative recombination
processes have to be avoided, which involves the elimination of the corre-
sponding deep centers. On the other hand, if one is interested in having a
short lifetime for non-equilibrium charge carriers, as in the case of fast tran-
sistors and photodetectors, one should intentionally introduce deep centers
in order to enhance non-radiative recombination. In general, deep centers
play a role of similar importance to that of shallow ones for semiconductor
devices, although in a completely different way.
Having discussed the general character and importance of deep centers,
we now will treat their electronic structure. In subsection 3.5.2 we will de-
velop a simple model of a deep center, the so-called defect-molecule model.
In subsection 3.5.3 we treat some methods of solution of the one-electron
Schrodinger equation for deep centers. The consequences of many-electron
effects in the electronic structure of deep centers will be discussed in subsec-
tion 3.5.4. In section 3.5.5 we display experimental and theoretical results
3.5. Deep levels 285

for selected deep centers, among them the vacancy in Si. the substitutional
impurity at,oms of the main groups of the periodic table. the group of 3d-
transition metals. and the group of rare earths. Also, we discuss the D X -
center and the EL2-center in GaAs.

3.5.2 Defect molecule model


The 'Tight 3inding' (TB) method developed in section 2.6 represents one
of the various procedures for calculating band structures of ideal crystals.
Unlike other methods it uses basis functions to represent the Hamiltonian
which are localized on the atomic length scale. Since the perturbation poten-
tials of deep centers are localized on the same scale, the TB method should
be particulady well suited for such centers. Of course, one must chose the
Hamiltonian matrix elements empirically in order t o arrive at useful practi-
cal results, and the results also cannot be expected to be very accurate in
a quantitative sense. However, the method should be suited to the deriva-
tion of simple models that exhibit the essential physical features of real deep
centers. The simplest among these models is the so-called defect-molecule
model, which we will introduce below. In doing so, many particle effects
and lattice relaxation will be ignored. We will mainly use the model to
demonstrate the existence of deep levels and to explore the symmetry of the
pertinent eigenfunctions.
At the outset we have to clarify which of the various tight binding basis
sets should be used for the representation of the Hamiltonian in the case of
a deep center. En the ideal crystal case considered in section 2.6,the atomic
orbitals or hybrid orbitals IhtRj) were not used directly, but rather.
we employed wavevector-dependent Bloch sums lukj) or ihtkj) formed from
them. This was advantageous because the translation symmetry of the crys-
tal could be exploited in this way. The latter symmetry no longer exists in
a crystal with a deep center. so that localized basis functions, i.e. atomic
or hybrid orbitals. can also be used without any loss. We select hybrid
orbitals because these produce drastic simplifications of the Hamiltonian
matrix which, like the Bloch sums of hybrid orbitals in the case of an ideal
crystal, allow the eigenvahes of the Hamiltonian to be calculated in closed
analytical form.
We introduce the defect molecule model using the vacancy as an example.
Later, we will apply this model to substitutional impurities from the main
groups of the periodic table. As the host crystal we take, in all cases, an
elemental semiconductor of group IV,like Si.

Vacancy

Figure 3.7 shows part of a Si crystal containing a vacancy. Symmetry consid-


286 Chapter 3. Electronic structure of semiconductor crystals with perturbations

erations make clear that the origin of the vacancy is not important, whether
it originated by removal of a Si atom from sublattice 1 or from sublattice
2. Here, we consider the removal of an atom from sublattice 1, and to be
still more specific, from the primitive unit cell at R = 0. The perturba-
tion potential V(x) is the negative of the potential produced by this atom
in the crystal. Because of the removal of the 1-atom, the hybrid orbitals
lht2Rt),t = 1 , 2 , 3 , 4 , of the four surrounding 2-atoms in the unit cells Rt,
pointing inwards, no longer have a hybrid orbital of a 1-atom to which they
can bind. They are called danglzng hybmds. The three other hybrid orbitals
at a surrounding 2-atom interact with inwardly directed hybrid orbitals at
atoms lying still further away (these atoms are not shown in Figure 3.7).
The hybrid orbitals at a pal ticular siirrounding 2-atom also interact among
themselves, including the one dangling hybrid at this atom. This means that
the dangling hybrids are coupled to the entire crystal through nearest neigh-
bor interactions. If interactions between hybrid orbitals at the same atom
are omitted, i.e. if the matrix element V1 introduced in section 2.6 is set to
zero, which means - cp, then the dangling hybrids are decouplcd from the
remainder of the crystal. They still interact only among themselves. Since
the atoms at which these hybrids are located are second-nearest neighbors,
this interaction is not within the framework of Erst-nearest neighbor inter-
action which we have been exclusively considering in the treatment of an
ideal crystal in section 2.6, but, rather, it ha6 the sense of second-nearest
neighbor interaction. The latter must be taken into account here in order to
arrive at non-trivial results.

Based on the approximations made above, the crystal with a vacancy


decomposes into two partial systems which do not interact with each other,
~ first, the partial system of the four interacting dangling hybrid orbitals,
one at each of the four atoms surrounding the vacancy, and second, the
yarlial system of all remaining hybrid orbitals of the crystal with the vacancy,
i.e. the hybrid orbitals of all atoms which are not directly adjacent to the
vacancy, and the three hybrids at each one of the four adjacent atoms which
&renot alrpady included in the first ptrrtial system. With some ambiguity
3.5. Deep levels 287

we can refer to the first partial system as a vacancy molecule: and to the
second the rest-of-crystal. This designation also encompasses h e term
tts
defect molecule model for the tight binding approximation described above.
For each hybrid of the rest-of-crystal another hybrid exists in the rest-
of-crystal which points to it. The Hamiltonian matrix elements between the
various pairs have the same value, namely the value given above in equation
(2.292) defining matrix element T/2. The energy spertriim of the rest-of-
crystal is therefore identical with t , h d of the infinite idea1 cryslal in bhe
simplest TB approximation? consisting of the bonding level t b = th - IVzl,
+
and the anti-bonding level E, = E,+ I&I. The splitting of these highly
degenerate levels into bands remains incomplete because of the neglect of
Ihe interaction between hybrids at. the same &om? i.e. the neglect. of r/l.
In calculating the energy spectrum of the first partial system, i.c., the
vacancy-molecule, we need the matrix elements of the pertarbed Hamilta-
nian H + V of the crystal with vacancy. The diagonal elements (ht2RtlH +
Vlht2Rt) are simply the hybrid energies ~h since the elements (h,t2Rt(V
lht2Rt) of the vacancy potential V between hybrid orbitals localized off the
%mancysit.e are small. One therefore has

In order to obtain the non-diitlgounel (second-nettrest neighbor) matrix el-


ements ( h t 2 R t l H f Vlht32FQ) between different dangling hybrid orbitals
t , t f t , one has to recaIl the symmetry of the perturbed Hamiltonian H +V.
Since, by creating a vacancy in sublattice 1) the two sublattices are no longer
eqiiivalent, the symmetry of the crystal is no longer given by thr full cubic
point group Oh, but rather by the tetrahedral group y d . Nevertheless, this
means that all non-diagonal elements are equal for symmetry reasons, such
that

wit,b W as second-nearest neighbor interaction energy. Because of the pre-


+
dominantly negative values of the operator fI V acting on hybrid orbitals
and the predominantly positive values of the hybrid orbital products. W
is expect.ed to be negative. Ihe absolute value of W must be determined
empirically.
The energy levels of the vacancy-molecule are obtained by diagonaliziug
the matrix
288 Chapter 3. Electronic structurc of scmiconductor crystals with perturbalions

Its eigenvalues Eyzz,4 read

(3.113)
(3.114)

IEY)=
Jz [IIZL~RI)
- IhaZRz)], (3.116)

1
IE,
VaC
) - -[ I ~ I ~ R
- IJh2R3)]
) I (3.117)
v5
~

1
138"")= - [lh12R1)- lh42R9)]. (3.118)
45
One can easily demonstrate that IEI") belongs to the irreducible represen-
tation A1 of the symmetry group Td of the vacancy, and the three functions
IE$uc), IEY'), IEiWc)to the irreducible representation T2 of this group (see
Table A.6 of Bppendix 4 ) . The eigenfunction of the A1-level resembles an
atomic s-orbital of a Si atom. and the three Tz-eigenfunctions are similar to
the three p-orbitals of such an atom. Evidently, the sp3-hybridization of the
atomic orbitals in a Si crystal is removed at a vacancy. The states are more
atom-like. in consequence of the fact that the crystal potential is no longer
fully effective in the vicinity of a vacancy.
In Figure 3.8. the e n e r e spectrum of the defect molecule model of a
vacancy is shown along with that of the rest-of-crystal. The A1-level lies
below the Tz-level because of the negative value of Vt-. Whether it lies
below or above the bonding level of the crystal depends on whether we
have 31W1 < l\Jzl or 31WI > lF'21. The question of whether it is found
in the valence band or in the energy gap, cannot be answered within the
defect molecule model because. therein, the valence band has shrunk to the
bonding energy level. It is just as difficult to decide whether the Tz-level lia
in the conduction bend or in the energy gap. Experiment and more exact
calculations, which we will discuss later in more detail. show that the -41-
level lies in the valence band. and the 7'2-level in the energy gap. With this,
3.5. Deep levels 289

Figure 3.8: Energy spectrum of the defect molecule model of ~ 1 .Si-vacancy along
with that. of the rest-of-crystal. The distribution of thc 4 electtona of the defect
molecule over the energy levels is also shown.

the T2-level is the actual deep level of the vacancy. Of the four electrons
of lhe neutral vacancy - each of the four dangling bonds yields one two ~

must he hosted by this lcvcl while the other two occupy the A1-level in the
valence band. IJsing the terminology introduced above, we may say that the
oxidation state of the neutral vacancy is V2+.
The defect moleciile model of the vacancy reflects the actual relation-
ships remarkably well. In any case it, provides a qualitatively correct phys-
ical picture of the electronic structure of a vacancy in group-IV elemental
semicondiictors. Refinements of this picture will be discussed further b e
low. Here we treat a second example to illustrate the defect molecule model
which emerges from the v.mancy by occupying its empty lattice site with an
impurity atom.

Substitutional impurity a t o m s w i t h sp3-bonding

We consider a substitutional impurity alom in an elemental semiconductor


of ~ o u IV,
p which we can again imagine as a Si rrystal (see Figure 3.9).
Let the siibstituted Si atom be that of sublattice 1 in the primitive unit cell
R = 0. Like Si, the impurity atom should belong to one of the main groups
11,111, IV, V, or VI of the periodic table so that the valence shell is formcd by
s- and p-orbitals, which all lie energetically higher than any other occupied
orbitals (in contrast to rare earth atoms). The perturbation potentials of
these impurity atoms in an elemental semiconductor of group I V possess, as
we know, both a short-range part and also a long-range Coulomb part. The
290 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.9: Defect molecule model of a sub-


stitutional main group &bonding impurity
atom in a tetrahedrally coordinated semicon-
ductor.

latter will be omitted from consideration below. This approximation does


not affect the answer to the question of whether a particular impurity atom
forms a deep level or not. although it influences the actual position of this
level in the energy gap. The latter cannot be determined within the defect
molecule model anyway.
The four sp3-hybrid orbitals of the impurity atom will be denoted by
IhtiO), and the four hybrid orbitals of the host crystal pointing in the
direction of this atom will be denoted by IhtzR,),t = 1,2.3,4. Xeglect-
ing interactions between the hybrid orbitals at the host atoms, the crystal
with a substitutional impurity decomposes, just as before in the vacancy
case, into two partial systems, first. the defect molecule with the 8 orbitals
jhtiO), jht2Rt),t = 1.2,3.4, and second. the rest-of-crystal with all remain-
ing orbitals. The energy spectrum of the rest-of-crystal again coincides with
that of the ideal crystal within the framework of the approximations used
here. The Hamiltonian matrix of the defect molecule is composed of elements
of the general form

fhtjRtIH + 17lht~jR~~)~ (3.119)


where j and j take the values i and 2 independently of each other. We
consider only the most important elements of this matrix, namely

In this, E ; and c i signify, respectively, the hybrid energies of the impu-


rity and host atoms, V2 corresponds to the matrix element of H of equation
(2.292) between hybrid orbitals at nearest-neighbor atoms of the ideal crystal
3.5. Deep levels 291

pointing toward each other, and W describes, as in the vacancy case, the sec-
ond nearest neighbor interaction between differing host atom hybrid orbitals
+
pointing toward the impurity atom. All elements ( h t j R t l H VlhtJRt,)not
listed above are neglected, among them also the Vl-like elements (htzRt(H +
Vlht,i&,) between different hybrid orbitals at the impurity atom, i.e. with
t f t. With these approximations, the Hamiltonian matrix of the defect
molecule may now be written down in explicit form. In doing so the rows
and columns are ordered in the sequence IhliRl), lh2iR2), lh3iR3), lh4iR4),
lh12R1)) lh22R2), lh32R3), I@&), and the Hamiltonian matrix is

L i 0 0 0 1720 0 0
0 e,O 0 0 vzo 0
0 0 6 i 0 0 0 v20
o a o a o v2 (3.124)
v2a o o ; W W W
0 v2o 0 W E k W W
0 0 v20 W W E k W
(0 0 0 I$ w w w e;.

This matrix has four distinct eigenvalues, two simple and two triply degen-
erate. The eigenfunctions of the two simple levels (we distinguish them by
indices b and a ) belong to the 1-dimensional irreducible representation A1 of
the tetrahedral group T d , and the two triply degenerate (again distinguished
by indices b and a ) belong to the 3-dimensional irreducible representation T2
of T d . The corresponding energy levels are, respectively, denoted as EiT/b
and E&Y:,b, whence

In Figure 3.10 these levels are plotted as functions of the difference ( c i - E);
between the hybrid energies of impurity and host atoms. This difference rep-
resents a measure of the strength of the short-range perturbation potential
of the impurity atom. In order to better understand the physical meaning of
the energy levels derived above, it is helpful to consider two limiting cases.
First, we assume the impurity atom to coincide with the host atom, which
means t i = c.; If one also neglects the second nearest-neighbor interaction
energy W , then the energy spectrum of the perturbed crystal of equations
(3.125) and (3.126) must coincide with that of the ideal host crystal in the
292 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.10: Energy levels Eimp


of a substitutional main group
sp3-bonding impurity atom in
a tetrahedral semiconductor as
function of the hybrid energy dif-
ference ( t i - e i ) . Horizontal lines
indicate the bonding level t b and
the anti-bonding level of the
host crystal.

simplest TB approximation, i.e. a bonding energy level q = ~ h IVzI,- and an


anti-bonding energy level E, = h + IV21 must emerge. This is in fact the case.
Thus it is also clear that the two energy levelsEiT,, of the perturbed crystal
correspond to bonding and anti-bonding states of A1-symmetry, and the two
energy levels E;Y$. to bonding and anti-bonding states of Tysymmetry. The
two undisturbed levels h f IV21 in Figure 3.10 encompass the energy gap of
the host crystal.
Second, we consider the limiting case of ( E ; - tk) tending toward $00
or -00 which, for a given host crystal, also means t i -+ +m or E ; -+ -00,
respectively. In the limit
E i + $00

equations (3.125) and (3.126) yield

(3.127)
(3.128)

and in the limit


Ei 4 --oo

we have
EiT = tk + 3W, (3.129)
3.5. Deep levels 293

zmp = Ezmp ~

(3.130)
EAlb Tzb - + --O0.

In the limiting case t i -+ +m, the two anti-bonding levels EiT and EkYZ
of (3.127) tend, along with t i , toward 00, i.e. leaving the energy spectrum
on its high energy side, while the two bonding levels EiY and EkYl limit
toward, respectively, the two Al- and Tz-levels of the vacancy of equations
(3.113) and (3.114). This is understandable because the limiting case EL ---f
+m means that no impurity hybrid energy level exists at finite energy and,
therefore, there is also effectively no impurity atom. In other words, there
is a vacancy.
With E ; -+ -m the two bonding energy levels EiY and E$T leave the
energy spectrum at its low energy side, while the anti-bonding levels limit
toward the two A l - and T2-levels of the vacancy of equations (3.113) and
(3.114). This occurs for the same reason as above in the limit t i -+ +m,
namely because t i + -00 means that there is a vacancy at the impurity
site.
For finite positive values of (EL - E;), the bonding Al- and Tz-levels can
occur in the energy gap, forming deep levels there, and for finite negative
values of (tL-~k) the same can happen with the two anti-bonding Al- and Tz-
levels. A look at Figure 3.10 shows that the most favorable candidate for an
impurity level to be in the energy gap is the bonding Tz-level if ( ~- 6;)i > 0
holds, and the anti-bonding A1-level if ( 6 ; - 6 ; ) < 0 holds. More rigorous
calculations confirm this conjecture in that they find exactly these levels to
be in the gap. Below, we will describe such calculations and discuss methods
for solving the one-electron Schrodinger equation (3.8) for the crystal with
a point perturbation.

3.5.3 Solution methods for the one-electron Schrodinger equa.


tion of a crystal with a point perturbation
The solution of the Schrodinger equation for a crystal with a point pertur-
bation begins with the determination of the effective one-electron potential
VPrt(x).In section 3.2 this task already was addressed in general terms.
Two remarkable differences were found in relation to an ideal crystal: First,
the effective one-electron potential VPert(x) depends on the electron popu-
lation of (localized) one-particle states, and second the atomic structure of
the perturbed crystal, which participates in the determination of the po-
tential Vwrt(x),is in many cases initially unknown and it must be self-
consistently calculated jointly with the electronic structure. Below, we will
eliminate such additional difficulties of electronic structure calculations for
point perturbations by assuming that the population of the center and the
atomic structure of the perturbed crystal are known. Thus the potential
Vwrt(x) of the center can also be considered to be known. In regard to the
294 Chapter 3. Electronic structure of semiconductor crystals with perturbations

electron-electron interaction contribution to V P d ( x ) , this only means that


the functional dependence of this contribution to VpeTt(x)on the eigenfunc-
tions of the one-electron Schrodinger equation is known, but not its actual
value. The form of this dependence is governed by the particular oneelectron
approximation used in the calculations - the possibilities include Hartree,
Hartree-Fock, Local-Density-Functional theory or the quasi-particle method.
Because of the dependence of the potential on the eigenfunctions of the
Schrodinger equation, any solution procedure must be iterated repeatedly
until self-consistency is achieved.
The three most important procedures are the cluster method, the super-
cell method and the Green's function method. As a rule, the replacement of
the real potential by a pseudopotential - a technique which has been \-cry
successful for ideal crystals (see section 2.5) - is also applied in the supercell
and Green's function methods.

Cluster method

Employing the cluster method, the infinite crystal with a point perturbation
is replaced by a finite part which contains the point perturbation at its cen-
ter and which. on the one hand, is large enough that the band structure of
the infinite crystal is almost completely devploprd and, on the other hand,
is small enough that the Schrodinger equation for it can be solved easily.
This finite part of the crystal is called a rlwter. The cluster with the point
pprturbatinn at its center is clearly just a large molecule. Its atomic and elec-
tronic struclureb can be calculated ubing the methods of quantum chemistry
for the determination of the structure o f molecules. In contrast to infinite
Lrystals, clusters have a surface. The latter also acts as a perturbation of
the periodic potential and can gi%erise to bound slates in the gap. Such
statrs must be remowd in a suitable way in order for the cluster method to
be applicable. To this end, different procedures h a w been proposed such as,
far example, the saturation of the dangling bonds at the surface by hydro-
gen atoms. Correspondingly, the surface levels are lowered deep below the
valence band edge and raised high above the conduction band edge, where
they cause no further difficulty.

Supercell method

At the boundary of a properly shaped ciuster, one can add an additional


cluster of the same kind. If one does this repeatedly and continues the
whole proress ad infiniturn, one finally arrives at an infinite crystal. The
primitive unit cell of this crystal is now, however, no longer that of the
crystal hosting the point perturbation. but it is the cluster, which in this
context is called a supeveell Thp crystal is referred to as a supercrystal. The
3.5. Deep levels 295

periodic repetition of the point perturbation in the supercell method causes


the discrete d w p levels to become k-dependent bands of h i t e widths. If
the supercell is made large enough, these widths are negligibly small and it
suffices to calculate the band structure of the supercrystal for one k-point
only, e.g., for k = 0. The appealing feature of the supercell method is that
the whole apparatus of band structure calculations for ideal crystals can
be exploited for the electronic and atomic structure determinations of deep
centers.
In the cluster and supercell methods one obtains the deep levels of the
perturbed crystal by a process in which one numerically calculates the energy
spectrum of a model system, i.e., of the cluster in the first method and of the
supercrystal in the second. Results for the band structure of the ideal crystal
are neither necessary nor useful in either method. Also, the decomposition
of the entire effective oneelectron potential into that of an ideal crystal
and a perturbation potential need not to be made. In the third method
for calculating the electronic structure of perturbed crystals, the Green's
function method, this decomposition is essential, and the band structure of
the ideal crystal is required.

Green's function method

The Green's function method employs techniques and insights of quantum


mechanical scattering theory. It provides results not only about bound states
with energy levels in the gap of the ideal crystal, but also on scattering states
having energies in the allowed energy spectrum of the ideal crystal. First we
consider the bound states.

Bound states: Kostcr-Slater method

To explain this method we write down the Schrdinger (3.21)equation for


the perturbed crystal once again in a somewhat modified form,

[E - HI+ - L"@. (3.131)


Since we are interested in bound states. the eigenvalues E which WP seek from
this equation lie outside of the bands, to be mare precise. in the iundamenlal
energy gap between the highest valence band and the lowest conduction
band. As tl preliminary we note that equation (3.131) can be formally solved
for @ by multiplying both sides by the inverse [E - H1-l of the operator
[ E - HI. The inverse operator [ F - H1-l stands in close relation t o the
Green's function of the unperturbed crystal. The retarded Green's finctaon
O " ( E ) is defined by the equation

1
GO@) = (3.132)
Eti6-H'
296 Chapter 3. Electronic structure of semiconductor crystals with perturbations

where 6 is an infinitesimal positive imaginary part which is set to zero after


serving to remove singularities at the energy eigenvalues of the unperturbed
crystal. This procedure assures that the wavefunction response of the un-
perturbed crystal conforms to the causality principle, in that the response
occurs only after the system has been perturbed. As our present interest
is in deep levels in the forbidden part of the energy spectrum of the ideal
crystal, not in changes of wavefunctions with energies in the already existing
continuous part, no singularities occur in G " ( E ) of equation (3.132) in the
range of E of interest to us, and we may ignore 6. Using G " ( E ) ,equation
(3.131) may be formally rewritten as

[@(E)V' - I]?) 10. (3.133)

Non-trivial solutions me possible only for energies E for which the deter-
minant of the matrix of the operator [ C 0 ( E ) V '- 11, in an appropriate or-
thonormalized basis set, vanishps, i.r. if

-
Uet[G"(E)V' 11 = 0 (3.134)

holds. The energy eigenvalues E satisfying this equation are deep levels.
These levels ran thus be determined by calculating the Green's function
G o ( E )of the unperturbed crystal and solving equation (3.134). This p r o w
dure is rcfrrred to as Koster Slater method, and equation (3.131) as Koster-
Slater equation
The Green's operator @ ( E ) can be cdriilattul if the band structure
E,(k) and the Bloch functions of the ideal crystal are known. The matrix
represenlation of d ' ( E ) in the Bloch basis reads

(3.135)

With this we only need to know the matrix elements (uk(V'1u'k') of the
perturbation potential V' in Bloch representation in order to determine the
deep levels using equation (3.134). However, i t would be more expcdicnt
for the matrix representation of the perturbation potential to use localized
wavefunctions as) for example, atomic orbitals. If one does so, then the
matrix representation of the (heen's opertrtoi is not as simple as in the Bloch
basis. A particular compromise in choosing a basis set for the representation
ol the eigenvalue equation (3.134) is the use of so-called Wunnzer finctzons.

Koster-Slater method in Wannier representation

Wannirr functions are linear combinations of all Bloch functions for a given
band index v of the form
3.5. Deep levels 297

(3.136)

Here the summation is over the whole first BZ,G is the number of primitive
unit, rdls in a ppriodicity rcgioa. and t k are certain phase a~glru.If the latter
are chosen properly, the Wunnier h c t i o n s IvRJturn out to be well localized
in the unit cell at the lattice point FL If the eigenvalue equation (3.134) i s
written in terms of this basis, it reads

The matrix elements of the Greens operator G o ( E )may be obtained by


means of (3.143) as

The matrix elements of the potential (vRIVjvR1 are particularly large if


R and R are identical, and both are equal to the lattice vector of the
cell which hosts the point perturbation. As before, we assume & = 0.
Neglecting all other elements we find

Employing (3.138) and (3.139)) equation (3.137) becomes

(3.140)

For the particular lattice point R = 0 it reads

{ 3.141)

Equation 13.141) forms a closed set of equations for the central cell compo-
nents ( ~ O ~ I Jof) $ only. If the latter are known, all other components (vRI$)
follow at once from equation (3.140). For a non-trivial solution of the ho-
mogeneous system (3.141) to exist, its determinant has to vanish separately,
i.e.
298 Chapter 3. Electronic structure of semiconductor crystals with perturbations

must hold. The diagonal elements (vOIGo(E)IvO) = G ! ( E ) of the Green's


operator in equation (3.142) may be obtained from (3.138) as

1 1
G;(E) =3
(vOlG*(E)IvO) (3.143)
G Ic E+ir-EE,(k)'

The dominant bands in equation (3.142) are those which form the gap, i.e.
the uppermost valence band u and the lowest conduction band c. If we
consider only them and neglect all others equation (3.142) becomes

or more explicitly

To solve this equation, the host Green's functions G:(E) and G : ( E ) , as well
the perturbation potential matrix elements, have t o be known. Below we
calculate the host Green's functions for isotropic parabolic bands of finite
widths. In expression (3.143) for G : ( E ) we rqlace the k-sum by an integraL
This yields

G 0J E ) = -- '
G 3 8x3 , 1 s t
d%i
~ ~E
1
-I- i t - E,(k) '
v = c, v. (3.146)

with st beiug the volume of a periodicity region. Introducing the identity


operator JrmdE'd(b" - E ) into this integral, G:(E:) may be written as

(3.147)

whrre p v ( E ) is the density of states per unit energy and spin state

R
p,(E) : - d3kd(E - &(k]) , v = c,w, (3.148)
h3 1.232
3.5. Deep levels 299

which differs from the DOS in equation (2.213) by a factor (0/2). In eval-
uating p,(E), we approximate the true valence and conduction bands by
isotropic parabolic ones, however, taking their band widths A E , to be the
same as those of the original bands. This corresponds to the use of an effec-
tive mass for each band averaged over the whole first B Z . The bandwidths
are introduced by putting the total number of states of the approximate
isotropic parabolic bands equal to G3. Then it follows that

(3.149)

3 G3
pc(E) = ~ ~ ~ ~- Eg)- O E, - AE,)]
[ B(E - ( E . (3.150)

where B(E) is the unit step function. We substitute these expressions into
the Greens function G ; ( E ) of equation (3.147). The E-integral is readily
done. The imaginary part ofG;(E) equals (-TI times pv(E). The real parts
ReG:(E) and ReG:(E) are given by

-I.;TIF] (3.152)

Although complex numbers appear in these expressions, they Eare in fact real.
This would hP obvious if we rcplaced the In-function by an arctan-function.
We avoid this because it is more convenient to handle the many-valued char-
acter of the In-function rather than that of the arctarz-function. Altogether,
the host Greens functions of OUT model depend on three parameters. the
energy gap EB, and the two band widths A& and AE,. The latter are
measure3 of the awrage kinetic energies of electrons in the vakncce or con-
duction bands large bandwidths mean large average kinetic energies (or
~

small eEective masses).


Later, the G r m s function method in Wannicr representation will be
used to address the question of whether or not main group impurity atoms
in tetrahedral semiconductors form deep levels in the gay.
300 Chapter 3. E1wtmnic Btructure of serniconductar crystals with perturbations

Scattering states

The perturbation potential also gives rise to changes within the energy bands
of the i d e a l rrystal. Of course, the w e r g i a of t h e w bands are still allowed
quantum mechanically in the presence of the perturbation, so that in this
rwpeet there is 110 change. However, change within the energy hands of the
ideal rrysttll IS indured by the perturbation in the form of a modified density
nf allowed energy levels pcr unit mnergy, i.e. in the form of a modifid density
ol statvs. To calculat? this change it is expedient to introduce the Green's
fiirirtion of the perturbed crystal,

1
(3.153)
G ( E ) - I; 4- i6 - [H+ V ' ].
According to formula (2.209). the imaginary part Im T r [ G ( E ) of
] the trace
represents (apart from a factor -1/ir) the density of states p ( E ) of the
syst,em, here thai of the crystal with the point perturbation. The Green's
function of the p e r t u r b 4 crystal oLrys the equation

G ( E )= G o ( E ) t G ' ( E ) V ' C ( E ) . (3.154)


which follows at once from the definition (3.1533 of G'(E). In quantum field
tbwry. this relation is known as the D p o n equatzon Using thk equation
and performing some simple calculations, the DOS expression (2.209) may
be brought into the form

1 d
p(E) - I m -In D e t [ G ( E ) ] . (3.155)
A dE
of the unperturbed crystal may be
In an analogous way? the DOS P O ( ! ? )
expressed in terms of the unperturbed Green's function Co(E). We seek the
change

M E ) = P(E)- P O ( W (3.156)
ol t h r nos clue l o the point perturbation. This change may be obtained
from (3.155) and (3.156) as

1 d
Ap(E) = --Irrt-lnDDet[l- G'(E)V'I. (3 157)
xdE
This relation can be used to ralculatc the total change of the DOS in the
entire enerRy range between -oo and t o o . Integrating A p ( E ) over this
interval, and considering the fact that G ' ( f i ) vanishes for li: -+ f m , yields

lm M

~ E A ~ (=.E0.) (3.158)
3.5. Deep levels 30 1

This relation signiEes that under the action of the perturbation potential,
the total number of states remains unchanged. This result already was used
in section 3.4 in lhe context of shallow levels, and there it was referred to
as Levznsons theorem. Here, this theorem means that for each new state of
the crystal created by the point perturbation, a state which existed without
perturbation must cease to exist. The deep states in the energy gap occur
therefore at the expense of band states. For each state occurring in the gap,
R stale is lost fiom a band,

dEAp(h:)- -
Lap. dBAp(E).

In the above derivat,ion of Levinsons theorem, no assumptions about the


(3.159)

spatial variation of the perturbation potential were made. This theorem,


therefore, also applies for purely long-range potenlials. Thus, we have proven
what was anticipated earlier in section 3.4, namely, that for each shallow level
in the gap, a state is lost from a band. In Chapter 4,where we will calculate
the electron population distribution over the energy levels in the gap and
the bands in thermodynamic equilibrium, this result will play an important
role.
At certain energy values in the bands, the DOS of the perturbed crys-
tal can display maxima or minima. One speaks of these as resonance and
antz-resonance state.9. These states emcrge when the short-range pertur-
bation polential can bind or anti-bind states at band energies. Since the
corresponding localized level is degenrrate with the band energy continuum,
the localization of resonance and anti-resonance states differs from that of
deep level states in the gap - with increasing distance from the center, the
eigenfunctions do not decay exponentially but oscillate with an amplitude
decaying to zero arcording to a power law.

3.5.4 Correlation effects


Correlation effects, as discussed in section 2.1, are important for the N -
electron system of a crystal with oneparticle states localized at a deep
centel. One of thcse e f k t s is based on tlie configuration dependence of
the Hartlee and exchange potentials. Another results from the fact that
Slater determinants, even thosc calrulatcd by means of configuration de-
pendent Hartrw and exchange potentials, are not exact eigenstates of the
N-electron Hamiltonian. The exact eigenstates are linear coinbinations of
different Sl&r determinants, an rffwt which is referred to as cor~fignralzon
znteraetzon (see section 2.1). The two correlation effects, configuration de-
pendenre and configuration interact ion, will be discussed helow for deep
centers. In regard to configuration dependence, we continue the general
discussion of section 3.2 here.
302 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Configuration dependence

For an ideal crystal, the energy eigenvalues E , of the onepartick Schrijdin-


ger equation have direct physical meaning. Apart from their sign, they are
the ionization energies I,, of the corresponding eigenststes Y , i.e. E , -I,.
In this regard, according to section 2.2, the ionization energy Iv is definned
as difference

(3.160)
of the total energy R t o ~ ( { v ) of) the N-electron system in the ionized stair
{v}, and the total energy E t o d ( { v } ]in theground state {v}. Theionized
state (v) differs from the ground state (v) in that a particle, which in the
ground state of the system occupies a oneparticle state of energy E,, is
transferred to a oneparticle state of energy 0 corresponding to the vacuum
level. Like in section 2.1 one says that a31 electron is removed from the
system. This expression has to he used with care, however, for if taken
literally i t misrepresents the charge neutrality of the system. The total
energy of the ground state follows, according to formula (2.541,by summing
all occupied oneparticle energies, followed by subtraction of the electrostatic
interaction energy of the electrons because the latter is doublecounted in
forming the sum of one-particle energies.
The equation E , = - I , is the content of Koopmans theorem, which
was explained in section 2.1. The essential requirement for the validity of
this theorem is the approximate population independence of the Hartree and
exchange potentials or, more generally, of the effective one-particle potential.
This requirement is not satisfied for a crystal with a point perturbation.
The potentials depend on the number n of electrons occupying oneparticle
states localized at the center, and so do the energy eigenvalues E , of the
one-particle Schrodinger equation of the crystal with a point perturbation.
We will denote these oneparticle energies by E:) henceforth to emphasize
this dependence explicitly. Of course, the definition (3.160) of the ionization
energy is also valid in this case, but it is no longer true that I,, represents the
negative oneparticle energy -E$.( That this cannot hold is immediately
clear if one recognizes that the eigenvalues IT?) and E P - l ) of the one-
particle Schrodinger equation with. respectively, n and n - 1 electrons at
the center differ from each other - the level EP- is deeper than the level
E p ) because the removal of an electron results in the positive core being
less strongly screened, so that the remaining electrons are more strongly
attracted. In this situation, we say that the electrons at the center relax on
the removal of an electron from the center. If one further considers that the
total energies of both systems enter in the definition (3.160) of I,, that of the
relaxed system with (n- 1)electrons at the center, and that of the unrelaxed
3.5. Deep levels 303

system with n electrons at the center, there is no way to explain why the
energy difference E b t d ( { v } v )- E ~ M ( { V } 'should) be equal to the negative
eigenvalue - E Z( ) of the unrelaxed system and not equal to the negative
eigenvalue -EP-') of the relaxed system. In reality, the ionization energy
E b t a l ( { v } v ) - E ~ M ( { v } ' )is not equal to either one, but lies somewhere in
between the two. It can be shown that it is approximately given by the
negative of the one-particle energy eigenvalue of the N-electron system with
the fictitious number ( n - 1/2) of electrons at the center. This is plausible
because the electron, during its removal, feels, so to speak, the potential with
n localized electrons half of the time, and feels the potential with ( n - 1)
localized electrons there also half of the time.
To carry out an exact calculation of the ionization energy of a center
which, in its ground state, has n localized electrons, the one-particle energy
eigenvalues E P ) are not sufficient. Applying equation (2.54),these one-
particle energies provide the total energy E b ~ ( { v } ' of ) the ground state,
but they cannot be used to calculate the total energy Etotal({v}v) of the
ionized state with n - 1 localized electrons at the center. To obtain the
latter, one also needs the one-particle energies EP-') of the N-electron
system with n - 1 localized electrons. In density functional theory, the two
total energies follow more directly: one determines the eigenfunctions of the
Kohn-Sham equation for the centers with n and n - 1 electrons, then forms
the corresponding ground state densities, and evaluates the total energy
functional (2.64) using these densities.
Ionization, i.e. exciting an electron to the vacuum level, is only one of the
various possible one-particle excitation processes of the N-electron system
of a crystal having a perturbation center. Generally, one may examine an
excited state { v } ~with ' ~ an electron in a formerly unoccupied one-particle
state Y' of energy below the vacuum level, and a hole in a formerly occupied
one-particle state v. The corresponding excitation energy Iuiv is given by
the total energy difference between the excited state { Y } ~ and ' ~ the ground
state {v}',

(3.16 1)
As in the case of ionization, the excitation energy Iulv is not equal to the
one-particle energy difference E,i - E,. Here, we are interested in excitation
processes involving changes of the populations of one-particle states local-
ized at the deep center. There are different types of such processes. Firstly,
in the final state, all electrons may still be localized at the center, but with
one electron having changed its localized one-particle state. Such excitations
are called internal transitions of the center. Secondly, an electron originally
localized at the center may undergo a transition to the bottom of the con-
duction band. This is referred to as a donor transition Thirdly, an electron
304 Chapler ,7. Electronic structure of semiconductor crystals with perturbations

Figure 3.11: Reletion between donor and acceptor ionazation levels.

from the t,op of thevalencr band may be transferred into a previously empt,y
one-particle state Iocalized at the center. This type of excitation is called
an acceptor transition. The excitation energy for an electron from the t,op
of the valence band to the bottom of the rnnduct,ion band defines the en-
ergy gap. As excitation energies of the N-electron syst.em of the perturbed
crystal, all transition energies discussed above may be plotted in the same
energy scheme, just as if they- were oneelectron levels. But. they are not:
they are one-particle ezcitation levels. More strictly speaking, one has donor
excitation levels E$V - E , - and acceptor excitation lcvels ~2~ =1 ~ 1 ~ .
The donor excitation level for a center D in the neutral charge state D(O) is
denoted by n(O/+),and the mceplor excit,ation level for a center A in its
neutral charge state A(0) is denoted by A ( O / - ) . If the donor excitation level
D ( + / 2 + ) at t.he single positively chargcd donor center D ( + ) also lies in the
gap, one has a dovhly i u n i m b k , OT d o u h k donor. If only the excitation level
D ( O / + ) lies in the gap, the donor D ( 0 ) is simply ionizable, or a single donor.
In the general case, t,here are m d t i p k donors. An analogous terminology i s
used for acceptors. Viewing the hand edges as int.erna1vacuum Icvcls one
0ft.m refers to donor and acceptor transitions as donor and acceptor ion-
izut,iuab, and to & curraponding excitation energies as ionization lez~elx..
Below, we will use this terminology
Acceptor ionization lcvcls may be traced back to donor ionization levels.
In fact, ionizing a neutral acceptor center A(0) leaves this center in singly
negative charge state A ( - ) , and a hole appear8 at the valcncc hand edge.
If one subjects the centex A(-) E D ( - ) t o a donor transition D ( - / O ) then
the center returns t.o its neutral charge slate D ( 0 ) A(O), and an electron
appears at the conduction band edge. As a resdt of this t,wo-step ioniza-
tion process of the center A(O), an electron-bolc pair is excited while the
state of the center has not changed. The sum of the two ionizalion energies
A ( O / - ) and D ( - / O ) equals, therefore, the (minimum) excitation energy of
an electron-hole pair, namely, the gap energy E , (see Figure 3.11). Gener-
ally, for an x:cept.nr A ( Q ) in charge state &, one has

(3.162)

If one measures thc acceptor ionization level relative to the conduction b u d


3.5. Deep levels 305

edge instead of the valence band edge (as is done in equation (3.162)), thus
setting A(&/(& - 1))= Eg - A(Q/(& - l)),then it follows from (3.162)
that

This unified description of donor and acceptor excitation levels makes it


possible to decide in simple way whether a given center X represents a
donor, an acceptor, both a donor and an acceptor or none of them. One has
(a) a pure donor, when the ionization level X(Q/(Q +
1))lies in the gap
and, simultaneously, the ionization level X((Q - 1)/Q) does not;
(b) a pure acceptor, if X((Q - l ) / Q ) lie in the gap and, simultaneously,
+
X(Q/(Q 1))does not;
+
(c) if both levels X((Q - l ) / Q ) and X(Q/(Q 1)) are found in the gap, the
center is both a donor and an acceptor. One then calls it amphoterzc;
- if neither of the two levels X((Q - 1)/Q) and X(Q/(Q + 1))lies in the gap
the center is neither a donor nor an acceptor.
The difference between the acceptor ionization energy X ( ( Q - 1)/Q) and
+
the donor ionization energy X(Q/(Q 1))of an amphoteric center X(Q) is,
by definition, the Hubbard energy 1.J. One therefore has

Since, coninionly, the Hubbard energy U is positive, the acceptor ionization


level commonly lies highcr in thc gap then the donor ionization level.
The number of electrons bound at a center in thermodynamic equilibrium
depends on the position of the Fermi level with respect to the ionization
levels. From the outset it i s clear that the ionization level X(Q/(Q + 1))
also marks that position of the Fermi level at which the charge state of the
cenlei changes: if E p lies just above X(Q/(Q I l)),the charge state Q is
realized, if E F lies just below X(Q/(Q + +
l ) ) ,the charge state (Q 1) is
realized. h o r n this observation one may conclude that the charge state Q
occurs when the Fernii Pnergy lies above the ionization level X ( y / ( y t 1))
and simultaneously below the ionization level X((Q- l ) / Q ) , i.e. between the
+
two levels X(Q/(Q 1)) and X((Q - l ) / Q ) . Of course, for this conclusion
to be valid, both levels have to be located in the gap.

Configuration interaction

If there is degeneracy among the various oneparticle states of an N-electron


system, then there is also degeneracy between the Slater determinants formed
306 Chapter 3. Electronic structure of semiconductor crystals with perturbations

from one-particle states of this type. Here, we concentrate on those degener-


ate one-particle states which are localized at the deep center. The existence
of such degenerate states was demonstrated in subsection 3.5.2 - the vacancy
in Si exhibits a %fold degenerate state transforming according to the irre-
ducible representation T2 of the symmetry group T d of the vacancy, beside
a non-degenerate state belonging to the irreducible representation A l .
To analyze the consequences of configuration interaction, we use a sim-
ple model: a deep center with 3 degenerate localized one-particle states
of symmetry Tz, occupied by 2 of the N valence electrons of the crystal;
the remaining N - 2 electrons are in extended valence band states. The
extended electrons contribute to configuration interaction only little; we dis-
regard them in our further discussion. Neglecting spin-orbit interaction,
the Slater determinants for the two electrons of the deep center are prod-
ucts of coordinatedependent and spin-dependent wavefunctions. Since a
Slater determinant itself is antisymmetric with the respect to the exchange
of two electrons, there are two possibilities for the behavior of the two fac-
tors in regard to particle exchange: either the spin-dependent wavefunction
is antisymmetric and the coordinate-dependent wavefunction symmetric, or
the spin-dependent wavefunction is symmetric and the coordinate-dependent
wavefunction antisymmetric. From quantum mechanics it is known that, for
a system with two electrons, the first case applies if the total spin S is 0,
and the second case if S = 1 (other values of the total spin are not allowed).
Using the three wavefunctions Iz), ly), 12) of the 7'2-state, one can combine
6 symmetric two-particle wavefunctions, namely

and 3 antisymmetric two-particle wavefunctions, namely

These are, altogether, 3' = 9, as one should expect.


If spin and exchange interaction are considered, the symmetric and an-
tisymmetric wavefunctions have slightly different energies because the ex-
change energy depends on total spin S , which differs for the two groups
of states. Within each group there is degeneracy, however, at least within
the framework of the one-particle approximation. If the configuration in-
teraction is taken into account, this degeneracy is removed. The result of
this removil can be derived by means of group theory, strictly speaking,
3.5. Deep levels 307

by means of the decomposition of symmetric and antisymmetric product


representations into irreducible parts (see Appendix A). For the symmetric
+ +
product one obtains (T2 x T Z }=~A1 E2 T2, and for the antisymmetric
product (2'2 x T2}a= T I . In these relations, the factors on the left-hand
sides are representations in oneparticle Hilbert space, while on the right-
hand side one has representations in two-particle Hilbert space. This means
that every state on the left-hand side can host 1electron, and every state on
the right-hand side 2 electrons. To avoid confusion, the representations in
oneparticle space are denoted by lower-case letters in this context, i.e. by
a l , a2,e, tl, t2 etc., instead by upper-case letters A l , A2,E , T I ,Tz etc., which
are common in group theory. Here, upper-case letters are used for two- (or,
generally, multi-) particle representations. This is chosen in accordance with
the notation for free atoms, where s,p, d , f represent oneparticle states, and
S, P, D , F represent many-particle states. Below, we will employ the distinc-
tion between lower- and upper-case representations wherever an ambiguity
might occur. As in the free atom case, the total spin S of the many-electron
+
state is indicated by the spin-multiplicity 2s 1, appended to the represen-
tation letter at its upper left. Using the notations introduced above, we may
conclude the analysis of our model deep center by stating that configuration
interaction will split its {t;}-configuration into the 4 different two-electron
energy levels 'A1, ' E , 'T2,and 3T1.
If more than two electrons are localized at the center, and if oneparticle
states of different irreducible representations are involved as, for example, in
the ground state configuration {aSt;} of the neutralvacancy with 4 electrons,
then the corresponding many-electron levels can be obtained in the same
way as above. The only difference is that the symmetric and antisymmetric
products to be decomposed into irreducible parts have more than two factors
and the factors are not necessarily the same.
The splitting of the many-particle levels of deep centers of crystals has
its counterpart in the fine structure of the many-particle energy levels of free
atoms with more than one electron. For such atoms, many-particle states,
formed from one-particle states of given total spin S and orbital angular mo-
mentum L , having different total angular momenta J , give rise to slightly
different energy levels. (Recall that the irreducible representations of the full
rotation group, into which the products of the irreducible one-particle repre-
sentations decompose, are distinguished by J . ) This fine structure splitting
has its largest effect for electron shells which are strongly localized, i.e. for
d- and f-shells (as opposed to s- and p-shells). One may expect, therefore,
that impurity atoms with unoccupied d- and f-shells, i.e. atoms of tran-
sition groups of the periodic table, should result in deep levels exhibiting
pronounced fine structure splittings. This is in fact the case, as we will see
below.
308 Chapter 3. Electronic structure of semiconductor crystak with perturbations

3.5.5 Results for selected deep centers

Below we discuss the structure of several deep centers which are important
either from the scientific or technological point of view. Knowledge about
these centers is, in every, case the product of combined experimental and
theoretical investigations. Since we have thus far treated only theoretical
methods, we first present a short overview of the experimental methods.

Experimental methods

One can divide the experimental methods for investigation of deep centers
into two groups, on the one hand, methods which measure ground state
properties of the centers, and on the other hand, methods which give exper-
imental data on center properties in thermally or optically excited states.
Among the methods of the first kind are measurements of magnetic
properties, like Electron-Paramagnetic Resonance (EPR);Electron-Nuclear-
Double Resonance (ENDOR) and magnetic susceptibility. These methods
provide data concerning the total spin S of the centers and, if anisotropy
effects are measured, also spatial symmetries. The chemical identity of the
centers can be determined (in addition to other methods) by means of mass
spectroscopy or of Rutherford Backscattering (RBS). Measurements of the
Extended-State X-ray Absorption Fine Structure (EXSAFS) provide data
about the geometrical ordering of the atoms in the vicinity of a point per-
turbation.
To investigate the excitation properties of deep centers, optical and elec-
trical methods are available. Ionization energies can be determined by means
of optical absorption spectroscopy and photoconductivity measurements,
mainly in the infrared spectral region. The cross-sections of deep centers
for emission of free charge carriers can be determined by means of time
resolved current or capacitance measurements at pn-junctions or at other
depletion layers. By suddenly applying a reverse bias at such a junction, the
deep levels are lifted relative to the Fermi level The new equilibrium state
of the junction corresponds to fewer electrons in the deep levels than previ-
ously. This state does not occur suddenly, however. it adjusts exponentially
through emission of electrons from the deep levels into the conduction band
(we assume deep donor levels here, the case of deep acceptor levels may be
treated analogously). In the conduction band, the electrons are freely mobile
carriers and are immediately sucked up by the positive electrode at the n-
region. This results in an exponentially decaying current, which for its part
leads to an increasing positive charging and. thus. an increasing capacitance
of the junction. The decay time of the current and the rise time of the asso-
ciated capacitance change are determined by the emission probability from
3.5. I?wp let& 309

the dttep centers. Measuring the capacitance rise time yields experimental
values for the emission probability. Of particular importatice is the so-called
Deep Level Trawaerat Spectroscopy (DLTS), wherein a reverse biased pn-
junction is exposed to periodically repeted voltage pulses of forward {deep
level filling) polarity. The recovery time for the capacitance change after a
filling pulse has been switched off, is measured as a fimrtion of temperature.
This fiinction exhibits maxima which, under certain conditions, can be used
to determine the ionization energies of the deep centers. Ionization energy
values from LILTS measurements and other thermal quililrrium technique
are, as a rule, smaller than optically measured values: the lattice has time to
rrlax if ionization proceeds thermally, and this lowers thc enerm of the finel
stat?. Optical ionization occurs instantaneously, hence lattice relaxation is
is not possible.
Perhaps the best understood point perturbation is the vacancy in Si, so
we initiate our discussion of particular deep centers with it.

Vacaricy in Si

The defwt molerule model of the vscancy was treated in section 3.2. It
predicts the existence of two bound one-particle states, a non-degenerate
al-state, and a tripiy degenerate t2-state. More rigorous calculations within
one-particle approximation (3araff, Schliiter, 1980; Bernholc, Pantelides.
1980) show that the al-level can be excluded as a deep state in the gap
because it lies in the valence band (see Figure 3.12a). As such, it is fully
occupied having two electrons. The t 2 -level can lie in the gap depend-
ing on how many electrons it hosts. ,4t maximum, this can be 6, and
at minimum 0. Thus there exist 7 charge states of the vacancy, namely
t7(2+),I,(+), i(O), I,(-), V(2-), Ir(3-) and V(4-). The one-particle en-
ergies of these centers differ by the Hubbard energy U. A simple estimate
yields 0.3 el for C3. Thus, many-body effects, more strictly speaking. con-
figuration dependencies of one-particle energies: should be important in the
case of the Si-vacancy. Other many-body effects, including configuration
mixing, are small, and a description of the vacancy in terms of one-particle
states, albeit configuration dependent ones. is approximately justified.
Owing to the value of about 0.3 e V for U , it may be expected that three
or four ionization levels codd fit in the Si gap of about 1.1e V . A4ctualIy,the
donor levels 5(+/2+), t(O/+), and the acceptor levels V ( - / 0 ) and V ( 2 -
/-1 are found there (as above the acceptor levels are counted relative to the
conduction band edge). In Figure 3.12 (part b). calculated positions of these
levels are shown. These are not yet final positions because lattice relaxation
has not yet been considered. The latter leads to changes which are discussed
below (see Figure 3.13). In charge state V(Z+), no electrons are available
310 Chapter 3. Electronic structure of semiconductor crystals with perturbations

for the population of the t 2 level, therefore no Jahn-Teller distortion of the


vacancy occurs. In charge state V ( + ) , the tz-level is occupied by 1 electron.
Its energy can be degraded through a tetragonal Jahn-Teller distortion. The
symmetry after the distortion is DM. For this group, the 3-dimensional
irreducible representation t 2 splits into the 1-dimensional representation 62
and the 2-dimensional representation c. The bz-level lies energetically below
the e-level and is single occupied. In charge state V ( 0 )the additional electron
can also be hosted by the bz-level. In order to gain additional energy, the
tetragonal Jahn-Teller distortion is strengthened. In charge state V (-),
population of the e-level begins. Through a further distortion of the vacancy,
which reduces its symmetry from D M to CzVithe e-level splits into two levels
and the additional electron is placed in the deeper of the two. This level can
also still host the additional electron of the charge state V(2-), with an
increase of the C2,-distortion. If one takes account of the energy shifts diie
to Jahn-Teller distortion, the resulting level positions are as shown in Figure
3 . 1 2 ~ .'I'he exchange interaction evidently plays only a minor role, so that
practically no spin-splitting of the vacancy levels occurs. Considering that
the wavefunction of the deep vacancy levels extend as far as the nearest
neighbor atoms, this is understandable.
A surprising result concerning the level positions in Figure 3.12~is that
the donor level V(O/+) lies below the donor Ievel V ( + / 2 + ) . The usual
ordering of the ionization level of the more negative charge state above the
ionization level of the less negative charge state is thus reversed. Formally.
it seems as if the Hubbard energy U would be negative instead of positive
for the transition V ( + / 2 + ) . One therefore also calls the vacancy in Si a
negatzwe-U center. Of course, the interaction energy between two electrons
at the center does not really change its sign, but the increase of ionization
energy due to the Jahn-Teller effect on the V ( + / 2 + ) transition amounts to
only about halfof that for the V(O/l+) transition since the Jab-Teller effect
is absent at the V ( 2 + ) center. If the vacancy was initially in the neutral
state, i.e. with Fermi energy lying above the V ( O / + )donor level and below
the V ( - / O ) acceptor level, with subsequent lowering of the Fermi energy
below the V ( O / + ) level, then the vacancy will initially capture a hole from
the valence band and pass into charge state IT(+), and from there, without
further change of the Fermi energy, i.e. spontaneously, it will capture another
hole passing into charge state V(2+).
The occurrence of an effectively negatke iJ at the vacancy in Si was first
predicted theoretically (BarrafF. Schliiter, 1980) and later found experimen-
tally in correlated EPR and DLTS measurements. This phenomenon has
now been observed at a number of other deep centers.
3.5. Deep levels 311

Figure 3.12: Deep levels of avacsncy in Si: a) Calculated ionization levels without
Hubbard corredions and lattice relaxation. b) Levels of a) with Hubbard correc-
tions. c ) Fsxperimetital ionization levels which, in addition, include Jahn-Teller
shifts. The numbers give the level distances from the valence band edge in eV.
(After Watkins, l S S 4 . )

Main group impurity atoms in tetrahedral semiconductors

The elements of the main groups of the periodic table are distinguished by
the fact that their valence shells arise born 3- and p-orbitals. The group-
IV elements, as well as thc IV-IV,111-VI.and 11-VI binary compounds,
form crystals with, respectively, diamond and zincblende structure in most
caws. Relow we refer to them as 'tetrahedral semiconductors'. If main
group elements appear as impurity atoms in such crystals, then they are
chemically bound in a manner similar to that of the hust crystal atoms which
they replace, i.e. through sp'-hybrids. That implies that the incorporation
of impurity atoms should be mainly substitntional in such crystals. In the
case of a compound semiconductor, the lattice site which is prpferred is the
one that belongs to the chemically most similar atom of the host crystal,
therefore the anion site, if the impurity atom is a nan-metal atom. and. the
cation site, if it is a metal atom. The solubility of the main group elements in
tetrahedral senlicntiductors is, accordingly, high. It lies between 10"
for chemically very similar systems such as GaAs.Al, and l O I 5 ~ r n -for ~
312 Chapter 3. Electronic structure of semiconductor crystals with perturbations

cl v

D2d - Dzd

d) v-
-o+d
-btc

Figurr 3.13: Defect molecule model of a vacancy in Si. Different charge states of the
vacancy are shown, taking into account the torresponding Jtthn-Teller distortions.
On the right-hand side, the basis functions of the irreducible representations of the
deep k v e l ~are indicated (a, b, c Bnd d bband for the dangling hybrids of the 4
surrounding atom). (After Watktns, l U U 4 . )

relatively different systems as, for inst ance, Si:Hg.

The perturbation polrntial of a main group impurity atom in a tetra-


hedral semiconductor contains, besides the short-range part, as a rule, also
a long-range Coulomb part. For isocoric impurity atoms, i.e. atoms whose
cores do not deviate too strongly from that of the host atom, the Coulomb
potential is in general the main contribution. As shown in section 3.4, this
potential leads to shallow donor or acceptor levels. These are the levels
principally involved if main group elements are used as doping atoms for
3.5. Deep ler7ek 313

Table 3.6: Experimental deep l e d positions of neutral main group substitutional


impurity atoms in Si (in eV). "Dash" indicates that no deep level occurs in the gap
or this particular substitutional impurity, meaning either that no localized state
exists at all (C, Ge, Sn, Pb), or that such states exist but are shallow (B, Al, Ga, P,
As, Sb. Bi). An empty space means that the correspondingimpurity atom is either
not incorporated substitutionally, or that there is no unambiguous experimental
data. The neutral vacancy (Vac.) is shown for comparison. [Data compiled from
Landoldt-Barnstein, 1982.)

?r
Be B C N 0
- - Ec - 0.14
hk A1 Vac. P S
- EC - 0.43 - Ec - 0.31
Zn Ga Ge As Se
+
E1, 0.32 - - - E, - 0.3
Cd In Sn Sb Te
E, + 0.55 E, t 0.15 - - Ec - 0.20
Hg T1 Pb Bi Po
1
Ec - 0.31 E, + 0.25 ~ -

tetrahedral semiconductors. For non-isocoric elements of the main groups,


however, the short-range potential dominates in general, and, in particular
cases. gives rise to deep levels in the gap (see Table 3.6). If an impurity
atom belongs to the same column of the periodic table as the host atom, one
says that the two atoms are isovcalent. In this case. the Coulomb potential
vanishes completely and the short-range potential remains as the only po-
tential contribution. Isovalent substitutional impurity atoms therefore lead
either to deep levels (this occurs if the isovalent host and impurity atoms are
chemically dissimilar, as in the case of GaP:P; or ZnTe:O), or no localized
levels OCCLU at all (this takes place for chemically very similar isovalent host
and impurity atoms as in the case of Si:Ge. which forms an alloy).

An important theoretical problem which has yet to be solved for main-


group impurity atoms in tetrahedral semiconductors is to understand why
certain elements cause deep levels in the gap while others do not. The impu-
rity problem mentioned was treated above within the defect molecule modeL
Although group-IV elemental semiconductors were assumed as host crystals,
the general results obtained there also apply to compound semiconductors.
3 14 Chapter 3. Electronic structure of semiconductor crystah with perturbations

Accordingly, a substitutional sp3-bonding impurity atom in a sp3-bonding


host crystal will introduce two bonding levels EL? and E t T of, respec-
tively, a1 and t 2 symmetry, and two anti-bonding levels E:T and E:; of
these symmetries (see Figure 3.10). Which of these levels are located in the
gap, and which are not, cannot be decided by means of the defect molecule
model. On the assumption that there is a deep level in the gap. one can
guess its symmetry as folIows: The perturbation potential V = e t - E: has
negative values (or is attracting) if the hybrid energy of the impurity atom is
lower than that of the substituted host atom, i.e. if the group number of the
impurity atom is higher than that of the host atom. Furthermore, V has
positive values (or is repelling) if the hybrid energy of the impurity atom is
larger than that of the host atom, i.e. for impurity atoms with lower group
numbers than the host atom. If, by varying the impurity atom, the per-
ER
turbation potential V= - E: takes increasingly negative values starting
from zero, then a deep level in the gap will evolve from the conduction band
below a certain negative threshold value. If the perturbation potential V
takes increasingly positive values starting from zero, then one expects a deep
level to arise from the valence band above a certain positive threshold value.
Using this observation, a look at. Figure 3.10 indicates that for negative (at-
tracting) perturbation potentials, the deep level should be the anti-bonding
u1-leve1, and for positive (repelling) perturbation potentials, the deep level
in the gap should be the bonding ta-level.

The above conclusions are essentially confirmed by the more accurate


Greens function tight binding calculations performed by Hjalmarson, Vogl,
Wolford, and Dow (1980). Some results of these authors. concerning al-
levels, are depicted in Figure 3.14. Since the a1-levels arise mainly from
atomic s-orbitals, they are plotted against the s-orbital energy of the impu-
rity atoms in Figure 3.14. Concerning the question of whether or not a main
group impurity atom gives rise to a deep level, the indications of Figure 3.14
agree surprisingly well with experiment. For the particular case of Si this
can be seen by means of comparison of Figure 3.14 with Table 3.6, which
summarizes experimental results for this host crystal. The data from both
sources agree that no deep levels exist in the case of the isovalent group-IV
atoms C, Ge, Sn, Pb. Among the group-V substitutional impurity atoms
N, P, ils, Sb, Bi, only IY gives rise to a deep level while all others result in
shallow levels (even those do not exist for the isovalent group-IV atoms).
In the case of group-\I atoms, Figure 3.14 indicates deep levels for substi-
tutional 0, S, and Se. while experimentally one also finds a deep level for
substitutional Te. For impurity elements left of column IV of the periodic
table, like Ga or Zn. the perturbation potential is positive, and the deep
level in the gap is expected to be tz-like rather than al-like. Such levels are
not well described in the TB calculations quoted above.
3.5. Dcep levels 315

>"
QI
Y -1
W

-2 L I I I I I
11 I I!
I
0

Eandedge
t -
A -1
>
Q)
u
w
-2 -
Cation Site
-3 I I I 1 I I A

-30 -20 -10 0


s - Orbital Energy (ev)
Figure 3.14: Deep levels for main group substitutional impurity atoms of various
tetrahedral semiconductors. The anti-bonding levels of al-symmetry are shown as
a function of the s-orbital energy of the impurity atoms. For further discussion, see
the main text. (After Hjalmarson, Vogl, Wolford, and Dow, 1980)

The fact that the perturbation potential is negative (attracting) for the anti-
bonding al-level and positive (repelling) for the bonding &level allows an
important conclusion on the nature of the deep impurity states: in both
cases, the corresponding wavefunct ions have larger amplitudes at the sur-
rounding host atoms than at the impurity atom. In this sense these deep
impurity states are more host-like than impurity-like.
Quantitatively, the al-level positions in Figure 3.14 differ from the experi-
mental ones. This is to be expected because of the great simplifications made
in the calculations. many-body effects, particularly, configuration dependen-
cies of the one-electron levels, have been neglected, and lattice relaxation has
not been taken into account (Scherz and Scheffler, 1993). To get an idea how
these effects would modify the results, we show in Figure 3.15 the expected
occupancy of the deep oneelectron levels in the particular case of Si as host
crystal, using the level ordering suggested by the above-quoted calculations
316 Chapter 3. Electronic structure of semiconductor crystals with perturbations

d) el fI

Figure 3.15: Energy levels and their populations for sp3-bonding main group im-
purity atoms in elemental semiconductors of group IV,within the defect molecule
model. The partial illustrations a) to h) correspond to impurity a t o m of groups I
to VII as follows: a - V, b - \I, c - VII, d - HI, e - 11; f - I.

and the defect molecule model. Of the 9 electrons of a Sirgroup-1 atom


molecule (remember that only N results in a deep level in this case), 8 are
hosted by the four bonding states and the remaining electron occupies the
anti-bonding al-state in the gap. In a 5t:groupVI atom molecule with 10
electrons, 2 occupy the al-level, and in the 3t:groty-VI atom molecule with
11 electrons, in addition 1 electron has to be placed into the anti-bonding
tl-level. This state of the molecule will certainly not be stable. As in the va-
cancy case, a Jahn-Teller distortion will occur which removes the degeneracy
of the ta-level and allows for a lower total energy by occupying levels shifted
downwards. The defect molecules of group-111. -11 and -I atoms have. respec-
tively, 7, 6 and 5 electrons. Of them 2 electrons are hosted by the bonding
al-level in the valence band. There are, respectively. 5, 4, and 3 electrons
to be placed in the bonding tz-state which forms the deep le\Tel in this case.
One may also say that this level hosts, respectively, 1. 2, and 3 holes. Again,
a Jab-Teller distortion will occur which lowers the total energy. Qualitative
3.5. Deep levels 317

differences from this model occui for elements of the first main group which
have no occupied p-states but relatively shallow closed d-shells (Cu, Ag) or
d- and f-shells (Au). Foi these atoms, d-electrons participate in chemical
bonding with the host crystal. We will discuss this problem in more detail
in the context of the transition metal impuiity atoms.
The question of whether deep levels exist or not for a particular substi
tutional impurity atom, ran he treated arralytzrally, employing the Wannier
representalion of the Greens function derived above. As lhis calculation
provides further physical insight into the formation of deep levels, we will
discuss it hclow, agdin taking Si as the host crystal. The Chens function
in Wannier representation has already been determined above (see equa-
tions (3.158) and (3.159)). For the evaluation of the Koster-Slatcr equation
(3.152) we need the matrix elemeuts V,,,,V,, V ,of the perturbation poten-
tial V in this representation. To calculate them we use the results of the
TR approximation of scction 2.5. In the simplest veision of this upproxi-
mation, the ILamiltonian of the ideal crystal is diagonalized by the bonding
and anti-bonding Bloch functions Ibtk) and In&), t = 1 , 2 , 3 , 4 , of equations
(2.303) and (2.304). Thus, in an approximate sense, the bonding and anti-
bonding orbitals IbtR) and latR) are the Wannier functions of the eigenvalue
equation (3.152). We will use thmn to grt explicit expressions for V,,,,,V,,,
and V,. Fiist, we write down the matrix representation of the unperturbed
Hamiltonian H between bonding orbitals. It is given by

(btRIHI4R) fhGRR,
h (3.165)
where E; is the hybrid energy of a host atom. For the perturbed Hamiltonian
EZ fV, one has t h e same matrix elements for all R and R with the exception
of R R= 0. The latter elements are

1
(btOlH i-
VlbtO) - S(E; + EL), (3.166)

where EL is the hybrid energy of the impurity atom. Siiicc none of the matrix
elements of (3.165) and (3.166) depend on t , one may identify IbtO) with the
Wannier function of the uppermost valence band. Then, taking the difference
of equations (3.166) and (3.165), an expression for (vO1V(wo) G V,, follows
SS

vlf, - v o , vu s1( 6 th - Eh).


h (3.167)
The factor in (3.166) arises because the Wannier function is equally spread
out over the two atomic sites of a unit cell, while the substitution of a n atom
occurs at only one site.
Analogoiisly, matrix elementu of fI and IT I V between anti-bonding
orbitals may be used to obtain an expression for (cO(Vlc0)I_ v,,,and matrix
318 Chapter 3. Electronic structure of semiconductor crystals with perturbations

elements between bonding and anti-bonding orbitals yield the expression for
(wO1V'lcO) V,. It turns out that all elements are the same, i.e.

v, = v, = v,, = LG. (3.168)


We seek deep level solutions Et of the Koster-Slater equation (3.152)located
in the gap, i.e. with 0 < Et < EQ. For such energies, the imaginary parts of
G:(E'} and G,!(E) are zero: and equation (3.152) transforms into

Solutions of equation (3.169) within the gap do not exist for all possible
values of the 4 parameters entering, i.e. the gap energy Eg, the two band
widths AE,, A&, and the perturbation potential constant Vo. This is evi-
dent if one considers the particular case AE,,v >> Eg,which can be realized.
The arguments of the logarithmic functions in G : ( E ) and G:(E) are close
to 1 in thia case, and the logarithms themselves become negligibly small. If
one assumes, in addition, that A & = AEw,then

+
Re[G:(E) G:fE)] - (F)--&[,,/=-
. fi] (3.170)

With this exprwsion! the deep level condition (3.169) takes the form

(3.171)

It has solutions Et only in the energy gap 0 < Et < E,, within which Et has
to be restricted anyway because the deep level condition (3.170) is valid only
there. For negative Vo, one necessarily has Ef > Eg/2,and for positive Vo,
Et < E,/2. For Vi -t +m, El approaches the value E,/2, i.e. the deep level
is pinned at the midgap position. This corresponds to the pinning of & to
the vacancy level discussed in subsection 3.5.2. In order for a solution Et of
equation (3.171) to exist at all, lV01 must exceed a minimum value J V O ) ~ ~ ~
given by

(3.172)

This minimum value of (Vo(follows from equation (3.171) by identifying Et


with the lowest possible deep level position in the gap namely Et = 0 in the
case of positive Vo, and by identifying Et with the highest possible value of
Et namely E, in the case of negative VO.Considering equation 13.1721, the
solution of equation (3.171) may be written in the closed form
3.5. Deep levels 319

Table 3.7: Perturbation potential matrix elements Vo = (1/2)(~: - $) (in eV) for
substitutional impurity atoms of main groups in Si, calculated by means of Herman-
Skillman s- and p-orbital energies. The latter are partially reproduced in Table 2.2.
(After Herman and Skillman, 1963.)

B e B C N 0 F
1.57 0.08 -1.41 -3.04 -4.8 -6.71
Mg A1 Vac P s c l
2.18 1.05 8.28 -1.12 -2.31 -3.56
Zn Ga Ge As Se Br
1.82 0.88 -0.04 -0.99 -1.97 -2.98
Cd In Sn Sb Te I
1.91 1.14 0.35 -0.42 -1.22 -2.03
Hg TI Pb Bi Po At
1.87 1.17 0.47 -0.24 -0.96 -1.68

According to condition (3.172) for a deep level to be formed, the absolute


value of the potential and the energy gap must be relatively large, and the
bandwidth relatively small. That lVol needs to be large enough is obvious.
The bandwidth must be sufficiently small because it corresponds to the av-
erage kinetic energy, which militates against binding. The gap is the energy
range where the deep level has t o be placed, so it should not be too small.
EnIarging the gap and lowering the bandwidth increases the likelihood of
forming a deep leveL
The condition lVol > IV0lmifi may even serve as a quantitative guide, as
the following numerical example for impurity atoms in Si demonstrates. In
this case one has Eg = 1.1 eV, and AE, = A E , = 3.3 eV is a reasonable
choice. With these values, IVolmiplis 1.21 e V . The perturbation potentials
V-0 for impurity atoms of the main groups II to VI of the periodic table,
calculated by means of equation (3.167), are listed in Table 3.7. According
to our model, a deep level should exist for lV0l > 1.21 eV, and should not
exist for IVoi < 1.21 eV. Comparison with experimental results or deep
levels of substitutional impurity atoms in Si shown in Table 3.6 reveals that
320 Chapter 3. Electronic struetrue of semiconductor crystals with perturbations

the rritprion is correct in all caws. pxrept for In and TI, which are close to
the border line of our model but still on thp side where no deep level6 exist.
while experimentally surh 1wrls art- found.

Transition metal impurity atoms

The transition metal ('YM)-impurity atoms of the iron group, i.e. those
with closed ad-shells (see Table 3.2), are incorporated in Si crystals prdom-
inantly on interstitial sites of tetrahedral symmetry, while the 5d-TM atoms
prefer substitutional cation sites. The position of t h e 4d-TM atoms in Si
lies between the two, i.e. both interstitial and substitutional incvrpuretions
are observed. In III-V and 11-Vl semiconductors, all three groups of TM
atoms prefer substitutional incorporation on cation sites. The solubility of
most transition metal elements in Si is relatively small, lying in the range of
rm-3. Higher values in the range of 101frm.3arc? reached in tetrahedral
compound semiconductors, and for M n in GaAs the solubility reaches as high
as 10" ~ r n - ~R;In
. also is a transition metal which gives rise to alloys with
cerhizl 11-VI compound semicondudors, such as, for example, (Zn, b1n)Te
or (Cd.hln)Te. Semiconduct,or alloys exist also with Fe and Co.
In tetrehdral semiconductors, most of the TM atmns form deep levels.
Among them, the levels of 'I'M-elements with closed 3d-shells (the so-called
3d-TM cleinents) are by far the bestst known. Concerning 4d- and 5d-TM
imyurily atoms, it is understood that. deep levels also exist in their rasp
(Beeler, Andersm. Sdieflcr, 1985 and 1990; Alves, Leihe, 1986). This is not
surprising since the Thl elements are all chemically very similar. Here, wc
restrict our considerations to the 3d-TM atoms. Them are the elements Sc!
Ti, V, Cr, Mn, Fe, Go and Xi. Sometimes, the neighbors of Ni to the right
in the periodic table, CU and Zn? are also included, although, owing t.o their
respective 3d1'4s- and 3d1'4s2-confi~~ations. these are not in fact transition
metals. In tetrahedral semiconductor crystals, however, they behave simi-
larly. Special interest in the investigation of 3d-Thl atoms results mainly
from t,he fact that their deep levels play an import.ant part in electronic de-
vices. They can serve as recorrilrizlation centers through which the lifetime of
non-equilibrium carriers is shortened, or as capture centers for h e carriers
which partially compensate the effect of dopant atoms. These effects are
imclesirable in most caws, thus one must. try to minimize the contamination
of the devicw by 3d-'I'M atoms. Occasionally, 3d-TM impurities can elso be
useful. 'I'his happens, for example, in the case of Cr in GaAs which, due to
its compensating effect! makes the crystals serui-insulitting.
There are numcrouq experimental investigations on 3d-'l'h.I impurit.y &oms
in tehahedral semiconductors (interested readers are referr4 to the book by
Fkurov and Kikoin, 1994). Beside donor and acceptor transitions alsn in-
ternal transitions of the 3d-TM imp1irit.y atoms are observed. Below we
3.5. Deep levels 32 1

concentrate on the donor and acceptor transitions. Figure 3.16 contains


data related to this, as well as showing the ionization energies of the free
3d-TM ions. The latter are larger than the ionization energies of the 3d-Thl
atoms hosted by semiconductor crystals by more than one order of magni-
tude. The case of free atoms is also striking in regard to the large Iyariations
of ionization energies between different ionization states. In semiconductor
crystals these variations are almost two orders of magnitude smaller so that,
as in the case of the Iacancy, ionization levels of several charge states may be
found in the e n e r a gap. Considering the substantially stronger localization
of d-electrons and the much weaker screening of their interaction in compar-
ison with the valence electrons this is surprising and needs explanation. We
will return to this point later.
Theoretically. the donor and acceptor ionization levels of substitutional
3d-TM impurity atoms are now well understood. A simple oneparticle
model which is supported by ab initio calculations, is a defect molecule
with the 3d-TM atom on a substitutional cation site at its center (see Figure
3.17). In this model, the TM atom is represented by the five 3d-orbitals and
the one 4s-orbital of its valence shell, and the representation of the crystal
is embodied in the four sp3-hybrid orbitals of the four surrounding cation
atoms pointing towards the TM atom. The energies of these orbitals will be
denoted by Ed, eS and eh, respectively. Considering the tetrahedral sgmme
try of the impurity center,we decompose the four sp3-hybrid orbitals of the
four surrounding cation atonis pointing lowdrdb the ThI atom into a state
with A1-symmetry and three states Itah) with T2-symmetry. The s-orbits1
of tlir TM atom is deformed by the crystal into an orbital of A1-symmetry,
and hhe five d-states decompose into two states ).1 with E symmetry and
t h r w states It%) with Tz-symetry. The two 01-orbitals interact with each
othrr tand give rise to a bonding state dwp in the valenct. band and an anti-
bonding state deep in the conduction band. The two estates remain without
bonding, their energy levels are likely to be found either in the gap or in the
valence band. The two triply drgeiicratc 12 states Itah) and 1t2d) do interact
mutually. 'rhe corresponding interaction matrix elements are of the type
' j r and bym {see section 2.6). Neglecting bbP-type matrix elements, one
triply degrnertlte bonding level Eb nnd one triply degenerate anti-bonding
level Ea arise. For these two levels, the same formulas apply 8s those which
wmt' derived in section 2.6 for interacting hybrids at nearest neighbor atoms
of an ideal crystal. Denoting the Vpp,-type matrix elemen1 by V h d , we hwe

(3.174)
where
322 Chapter 3. Electronic structure of semiconductor crystah with perturbations

Figure 3.16: Experimental d'/do d%' d?d2 dyd' d%l' d?d5 d?da d%' d%lnd'o/d9
ionizatjon energjes of free 3d- I I I I I I I ' 1 11
Tzvl ions (lower part), as well
as donor and acceptor ioniza-
tion energiea of 3d-TM impu-
rity atoms and ions in various
host semiconductors (upper five
parts). (After Zernger, 1986.j

GoAs to/-
2.4

The corresponding thecomponent wavefunctions It2b) and Itza) of, respec-


tively, energy Eb and E,, are

where we set

1 s
a2 = - ( I - -
a)' a 2 = -1( l + - s) . (3.177)
2 A
3.5. Deep levels 323

c -band

t2a \ t2
4s
3d

v - band

Figure 3.17: Energy level diagram of the defect molecule model for a substitutional
3d-TM impurity atom in a tetrahedral semiconductor. Explanations are given in
the text.

Since the two bonding and anti-bonding eigenvectors Itzb) and Itza) of equa-
tion (3.175) are linear combinations of basis functions with tz-symmetry,
they also have t2-symmetry (as already indicated by the notation). It turns
out that the bonding tzb-level lies in the valence band, the non-bonding e-
level in the valence band or the gay, and the anti-bonding tk-level is in the
gap above the e-level or in the conduction band. In regard to deep levels of
the 3d-TM atom i s in the energy gap, therefore, the 2-fold degenerate e-level
and the abovelying triply degenerate anti-bonding tp,-level are candidates
(see Figure 3.17). The eigenfunctions of the e-level are linear combinations
of the &orbitals of the TM atom and are therefore strongly localized at the
latter. Considering localization of the t a b and tn,-leveh, the position of the
two orbital energies F d and t h relative to each other is decisive. If Ed lies
deeper than Ch, then a! > ,8 holds, meaning that the bonding tzb-level is
mainly formed from the d-ststes of the TM atom, while the anti-bonding
tza-level, which forms the deep level, arises to a considerable extent from the
tz-components of the four sp3-hybrid orbitalb of the surrounding host crys-
tal atoms. Exactly this picture emerges from the above-mentioned Greens
function ab initio calculations (Zunger, Lindfeldt, 1983). In Figure 3.18 the
calculated oneparticle energy levels are shown for some 3d-TM atoms in
Gap.
The most striking feature of the above description of the electronic struc-
ture of substitutional 3d-TM impurities is the existence of deep levels whose
wavefunctions are spread out over the surrounding host atoms. These levels
324 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.18: Calculated one-


electron energy levels of substi-
tutional 3d-TM impurity atoms
in Gap. (After Zunger and Lindt-
felt, 1983.)
4a
3

*\1-

-jFq
-2
-3

are host-like to a certain extent rather than purely TM atom-like. This pic-
ture stands in remarkable contrast to the so-called ionic model, which was
long believed to be correct for substitutional TM atoms also. In this model,
the deep level is strongly localized at the TM atom. Today there exists clear
experimental evidence that this is not the case. The ionic model is based
on the so-called ligand field theory. This entails an approach to the deep
level problem for impurity atoms which differs fundamentally from the one
used in this book. Within ligand field theory, it is assumed that an impurity
atom X in the crystal has essentially the same electronic structure as the
free X"+-ion, where V + means the oxidation state of the impurity atom in
the crystal introduced above. The energy levels and wavefunctions of this
ion are weakly disturbed by the crystal field at the impurity site, an effect
referred to as crystal field splitting. While this model applies relatively well
in the case of ionic crystals (where ch lies deeper than c d ) , it evidently fails
in the case of the covalent or partially covalent tetrahedral semiconductor
crystals. The oxidation states of ligand field theory are, however, also r e
produced in the approach taken here. To demonstrat,e this we consider the
example of a Co atom substituting the metal atom in a 111-V compound.
There are 9 electrons in the 3d74s2 valence shell of Co, and 5 in the valence
shell of the group V atom to which the Co atom binds. 8 of these 14 electrons
are hosted by bonding states of the deep center (2 by the bonding al-state,
and 6 by the bonding t 2 state). S i x electrons remain for the population of
the deep impurity states, i.e. the non-bonding e-states and the anti-bonding
t2-states. Since 9 electrons are expected at a neutral Co atom, the oxidation
state of Co in a 111-V compound is Co3+. Substitutional Co in a 11-VI com-
pound has the oxidation state Co2+ since one additional valence electron is
provided by the group VI atom of the host crystal.
3.5. Deep levels 325

3.0

-552 2.0
LT
W
z
w
1.o

0.0

Figure 3.19: Calculated inultiplet structure for substitutional 3d impurity atoms


in ZnS. [After Faatio, Caldas, and Zungcr, 2984)

Ligand field theory accounts for the fine structure of the deep impurity lev-
els, which is not explainrd by the approach taken above. Although this
approach provides qualitative understanding of deep 3d-TM centers in tetra-
hedral semiconductors, the excitation energies it yields are not correct quan-
titatively, particularly not for the internal transitions of the 3d-TM centers.
To be quantitatively correct, this approach must be refined by including
many-body effects, in particular, configuration interaction. While the latter
effect does not play an important role for main group impurity atoms, as we
have scen above, it becomes important for T M impurity atoms with their
strongly localized open d-shells (as was specultited in section 3.5.4 using the
analogy with frcc I'M atoms). Figure 3.19 demonstrates this in thc case of
TM atoms in ZnS.
The differences between the donor or acceptor ioniaation energies of a
particular TM atom in different host crystals exhibit interesting behavior.
Experimentally one finds that these diffeiences are almost independent of the
atom considered. Therefore the ' ionization-encrgy-versus-'I'M atom curves'
of different semiconductors are parallel t o each other and ran be translated to
overlay by rigid displacements along the energy axis (Caldas, Fazzio, Zunger,
1984; Langer, Heinrich, 1985). In Figure 3.20 this fact is shown for acceptor
326 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.20: (a) Experimental acceptor ionization energies of 3d-TM impurities


in 111-V semiconductors. (b) Experimental donor (open symbols) and acceptor
(filled symbols) ionization energies of 3d-TM atoms in IFVI semicoIiducton. (After
Langer and Heinrich, 1985.)

and donor ionization levels of a 3d-TM impurity atom in a series of I11 V and
11-VT compound semiconductors. The rigid displacement along the energy
axis has an important physical meaning as we will discuss further below.
Here, we will provide a simple explanation for the similarity of the 'ioni-
zation-energy-versus-TM atom' curves in different semiconductors (Tersoff,
Harrison, 1986). In this discussion we will use the defect molecule model
of 3d-TM impurity atoms presented above, in addition, however, the depen-
dence of the energy levels f d and Eh on their population will be taken into
account in a self-consistent way. Let n d , n h and n:, n: be the electron num-
bers in, respectively, the atomic d- and sp3-hybrid orbitals, the former after
charge transfer between the impurity atom and the crystal has taken place,
and the latter before. The dependence of the levels Ed and ~h on population
may, in the linear approximation of equation (3.25), be written as

Ed = t d0 + Ud(nd - r i 0d ) , Eh = 1,0 tJh(nh - nh).


0 (3.178)
Here c,: :E denote the energy levels without charge transfer, and [Jd, U h
are, respectively, the Hubbsrd energies of the d- and the h-states of the sp3-
hybrids. The electron numbers in the bonding and anti-bonding ta-states
will, respectively, be denoted by rib and n,, and the electron number for the
estates by ne. Of the total number of electrons of the defect molecule,

n d = nbu! 2 + nap2 i-
n, (3.179)
3.5. Deep levels 327

are in d-states, and

n h = n b P 2 +,,a2 (3.180)
are in h-states. The bonding T2-level Eb lies in the valence band, its three
states are therefore completely occupied and it holds n b = 6. The difference
6 = (Ed - h ) / 2 between the d- and h-states adjusts self-consistently: if Eh
increases, then 6 becomes small and, with this, also ,B and n h are small. As
a consequence of this, n d increases which, according to (3.178), results in an
increase of d and, hence, an increase of 6. Self-consistency is then achieved
if the two equations (3.178) for the level positions t d and Eh are satisfied with
values n d and n h from equations (3.179) and (3.180), respectively, and values
a 2 , P2 from (3.177). For the level difference 6 = ( d - ~ h ) / this
2 condition
yields

where 260 = (E!-


+uh

ci)
" ,(nu
1
+ 6 ) - -(nu
2
6
- 6)-A - n:)] (3.18 1)

denotes the level distance without charge transfer.


The crucial point at this juncture is that [ I d , because of strong localization
of the &electrons and wcak scrcening of their interaction, has very large
values (on the order of magnitude 10 e V ) . Therefore, the factor multiplying
U d in equation (3.181) must practically vanish itself in order to satisfy this
equation, leading to the approximate rclation

1
+ 6 ) + s1( n u- 6 ) A-6
i ( ~ a
0
1 ne - n d w 0, (3.182)

which uniquely determines 6. According to this relation, the d-level adjusts


in the crystal in such a way that the charge transfer between the crystal and
the d-shell practically vanishes (of course, it cannot vanish completely be-
cause adjustrrierii of the levels requiies a bit of rhargc to bc transferred). This
adjustment is called the 'neutrality level'. According to equation (3.181), the
self-consistent value of h and, thus, also the energy separation tId-h = 26 be-
tween the d- and < h level, is indepmrlent of Ch. Since t h is the only quantity
which changes substantially in the series of the 111-V and 11-VI-cornpound
semiconductors, the separation between Ed and Eh is the same in different
semiconductors. This also holds for the anti-bonding tza-level dnd for thc
non-bonding e-level, i.e. for the two levels which are candidatev fol deep
levels in the gap - also, their clihtance from c h is independent of Eh. For
the ta,-level this follows from equation (3.174), and for the e-level from the
fact that its position is determined by the location of the d-level in the c ~ y s -
tal. We thus arrive at the conclusion that the deep levels of 'I'M atoms in
328 Chapter 3. Electronic structure of semiconductor crystals wilh perturbations

the gap are pinned at the hybrid level F ? ~of the surrounding host anions.
If the coupling hetwccn the anion-hybrids and the rest-of-crystal is taken
into account, the anion-hybrid energy Eh has to be replaced by the average
cation-anion hybrid energy of the crystal.
The rebUlt8 above explain lhe experimental finding that the deeplevel-
versus-Thl atom-curves for two different semiconductors ran be tranbleted
to overlay each other by rigid displacements. Moreover. they associate the
magnitude of the displa~ementswith the difference between the average hy-
brid energies of the two semiconductors. Thls quantity has a close relation
to the valence band discontinuity at a semiconductor heterostnicture. as will
be discussed iu more delail in section 3.8
The large Hubbard energies of the atomic d-states of the IM impurity
atoms and thFir pinning at lhc dangling hybrid lev& of the surrounding
atoms are also contributo~yto the striking phenomenology noted above. in
that the deep levels of TM atom8 depend only relatively weakly on their
population, so that levels of several charge states can fit in the gap. In fact,
if an additional electron is put in the deep tzo-lwel, and, with this, also part
ofit IS added to the TM atom, the d-level of the latter will be shifted up and
the bondmg tB-level will be depolarized ( a becoming smaller and /3 larger
in equtliiori (3.177)). This means that electron charge density will flow from
the Thl atom to the surrounding host atoms. in almost the same measure as
was added t o the T M atom when the additional electron was placed in the
deep ant i-bonding &-level. Since the Hubbard energies of the s- and p-shells
of the host atoms are substantially smaller than those of the &orbitals of
the Thl atoms, the deep TM-levels are shifted up only by several tenths of
an e V rather than 10 el, for the case of the free TM-ion if one more electron
is p l a ~ e dat the center (Haldan, Anderson, 1976).

Cu, Ag, Au in Si

These three elements (abbreviated below as Nhl, for noble metal) play
an important part in silicon device technology. As impurity atoms, they
constitute effective capture centers for non-equilibrium charge carriers. Their
high solubility in Si is remarkable (in the region of cmd3 to ~ r n - ~
at melting temperature), as is their fast diffusion. Their incorporation in
the host crystal is predominantly substitutional. The elements tend to form
complexes with other elements. for example, Au with 0. Fe or 3d-TMs. This
makes it somewhat difficult to identity the pure substitutional NM centers.
Although the elements Cu, Ag and Au belong to the fast main group of
the periodic table, their behavior its impurity atoms in tetrahedral semicun-
ductors resembles that of transition metals. This is understandable if one
considers that for these elements only the outer s-shell, but not the outer
p-shell, is occupied, and that the closed d-sheIl (Cu, Ag) or d - and f-shells
3.5. Deep levels 32 9

(Auj lie energetically relatively shallow below the s- and p-shells. According
to calculations by Fazzio, Caldas and Zunger (1985). the anti-bonding t 2 -
state lies in the energy gap, just as in the case of transition metals, and as in
the latter case, this state arises from the interaction of the impurity-derived
d-states (more exactly their tg-component) with the tg-components of the
dangling bonds of the surrounding host atoms. Because of the relatively
weak localization of this deep impurity state, its Hubbard energy is small
so that. again. several ionization levels can fit within the gap. For all three
K M atoms, the calculations result in a Nh/l(O/+) donor level, and a NM(-/O)
acceptor level. In the case of Ag and Au, the amphoteric character has also
been conhmed experimentally.

Rare earth atoms

The rare earth (RE) atoms are among the following elements: Ce, Pr, Nd,
Pm, Sm, Eu. Ac. Th. Pa, U. Np, P u , Am, and Cm. The valence shell config-
uration of the majority of these atoms may be described as 4fn6s2 with the
number n of f-electrons varying from 2 for Ce to 14 for Yb (see Table 3.3).
Exceptions are Gd and Tb which have one 5d valence electron in addition.
The investigation of RE impurities in Si as well as in the 111-V and 11-VI
semiconductors has received new stimulus very recently. The reason for this
is the luminescence of RE impurity atoms in these crystals in the technologi-
cally interesting visible and infrared spectral regions. In equilibrium the RE
atoms are installed both substitutionally as well as interstitially. The equi-
librium solubility of the RE atoms is, however; rather small. For practical
applications, non-equilibrium incorporation techniques like ion implant ation
must be used, although, generally, only parts of the implanted atoms are
optically or electrically active. For Erbium (Er), impurity concentrations of
- ~ been reported in Si, and of about 10l8 C W I - ~ in Ga4s.
about lo1 ~ r n have
The technologically interesting luminescence discussed above is caused
by internal transitions within the 4f shell of the RE atoms, more strictly
speaking, between the various levels arising from the many-particle levels of
the 4f-shell under the influence of the crystal field. The 4f-orbitals remain,
in fact. almost unchanged by installation of the RE atoms in the crystal. This
may be traced back to the fact that these orbitals are strongly localized and,
moreover, strongly shielded by the 6s-electrons shutting out influences of the
surrounding crystal. Furthermore, for RE elements heavier than Nd, there
is also strong shielding by the completely occupied 5s- and 5p-shells lying
outside the 4f-shells for these elements. While internal electron transitions
within the 4f-shell have been studied already for a long time, transitions to
energy levels outside of the 4f-shell, including donor and acceptor transitions
into conduction and valence band states of the crystal, have been subjected
to more detailed investigation only recently.
330 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Ce Pr Nd Fm
' Sm Eu Gd Tb Dy Ho Er Tm Yb
8 ! ' I ' I j 1 1 1 I " ,

6
4
2
5 0
v

P -2
g-4 k E(3+/4+) I
-6
-8
-10
1I I
-12 , I ! I I / I I , ! I
I
Ce Pr Nd Rn Sm Eu Gd Tb Dy Ho Er Tm Yb

Figure 3.21: Donor ionization levels of substitutional rare earth impurity atoms in
different semiconductors. (After Delerue and Lannoo, 1994.)

A rough but qualitatively correct picture of the electronic structure of sub-


stitutional RE impurity atoms in tetrahedral semiconductors may again be
obtained by means of a defect molecule model (Delerue and Lannoo, 1991).
In this model, the 4f-orbitals neither interact with the 5d- and 6s-orbitals
of the RE atom nor do they couple with the orbitals of the surrounding host
atoms. The RE atom is represented by its 5d- and 6s-orbitals only. The 4f-
orbitals, nevertheless, are involved in forming deep states because electrons
are transferred to or from them. The host crystal is described by the four
sp3-hybrid orbitals at the four neighboring atoms pointing towards the RE
atom. The model is therefore largely analogous to that of 3d-TM atoms de-
scribed above: The five 5d-orbitals decompose into two e-orbitals and three
tz-orbitals, and the 6s-orbital becomes a al-orbital under the influence of
the tetrahedrally symmetric crystal field. The four sp3-hybrid orbitals at
the neighboring atoms pointing toward the RE atom split into one a l - and
three t2-orbitals. The interaction between the two al-orbitals results in a
bonding and an anti-bonding al-state lying deep in the valence and conduc-
tion bands, respectively. The e-states remain without bonding to the host
crystal, while the host and RE derived t2-states couple to each other forming
bonding and anti-bonding tz-states. The differences between 3d-TM and RE
impurities are essentially of a quantitative nature. The 5d-levels of the RE
atoms are higher in energy and therefore closer to the hybrid levels at the
neighboring atoms than are the 3d-levels in the 3d-TM case. Because of this,
the e-level and the anti-bonding tz-level of the RE atom (which may lie in
the gap in the case of 3d-TM atoms) are lifted into the conduction band.
3.5. Deep levels 33 1

The position of the If-shell with respect to the above levels still has to be
determined. In the model by Delerue and Lannoo this is done assuming that
the distance between the 4f-shell and the average 5d-level in the crystal is
the same as that between the 4f-shell and the 5d-shell in the free RE atom.
The ionization levels of the 4J-shell calculated by means of the above
defect molecule model are shown in Figure 3.21 for various RE atoms and
host crystals. Two diffcrcnt oxidation states of the RE atom, RE3+ and
RE2+, have been assumed. One recognizes that, for InP, the oxidation state
RE3+ is stabk in most rasrs, because the RE(3-t /4 t ) ionization levels are
still below the Fermi level, while the RE(2 t/3+) ionization levels are above
it. For CdTe, the oxidation state RE2+ is found to be stable in most rases.
With a few exceptions, these predictions are confirmed by experimental in-
vestigations. Oxidation states can also be derived in a more direct way from
the above defect molecule model. In a 111-V compound host crystal, the RE
+ + +
defect molecule has 2 n 5 electrons, 2 n of them from the HE atom, and
5 of them from the host crystal anion. These electrons orrupy either states
in the 4f-shell, or are in bonds between the RE atom and the crystal. The
bonding al- and t2-states host 2 f 6 = 8 electrons provided that they are
energetically lower than the f-shell, which in fact turns out to be the case.
+ +
Thus, 2 n 5 - 8 = JZ - 1 electrons remain for the f-shell. The oxidation
state of a neutral RE atom is RE3+ in such circumstances, because 3 elec-
trons are missing at the RE atom, the one f -electron and the two s-electrons
in bonds with the host crystal. In 11-VI compounds, the oxidation state of a
neutral RE atom should be RE2+ according to this consideration, since the
host crystal anion delivers 6 electrons instead of 5.
A striking feature of the ionization-energy-versus-RE-atom-curves for
different semiconductors in Figure 3.21 is that they can be brought to over-
lie each other by rigid relative displacements. This is reminiscent of the
ionization energies of the 3d-TM atoms where the same feature is observed.
The reason is similar to that of the 3d-TM case: one the one hand, the
4f -levels are pinned electrostatically to the average 5d-levels by means of
the self-consistent charge exchange between the d- and the f-shell with its
large Hubbard energy, and on the other hand, the 5d-levels are pinned at
the average hybrid energy levels.

As-antisite defect i n G a A s (GaAs: A s G ~ )

The interest in this defect arises mainly from the fact that it is closely con-
nected with the so-called EL2-center which is one of the most common point
perturbations in GaAs. The EL2-center is a double donor and plays an im-
portant role in GaAs electronics. Through deliberate use of its properties,
p-type GaAs, which normally arises in crystal growth, can be transformed
into semi-insulating GaAs which is required for the manufacture of GaAs
332 Chapter 3. Electronic structure of semiconductor crystals with perturbations

MESFETs and other electronic devices based on GaAs. However, in regard


to materials for light emitting optoelectronic devices, the concentration of
EL2-centers has to be kept as small as possible because EL2 acts to quench
luminescence. Whether the As-antisite defect GaAs: ASG, is really identical
with the EL2-center, or if it is only part of a complex which represents the
EL2-center, is still somewhat controversial today. Nevertheless, important
properties of the EL2-center , in particular its structural metastability, can
be explained in terms of the simple GaAs:Asc, antisite defect alone (Chadi,
1992, Dabrowski, Scheffler, 1992).
The GaAs:AsG, antisite defect may be viewed as a special case of a
substitutional main group impurity atom with sp3-bonding in a tetrahedral
semiconductor. For the latter point perturbation, the defect molecule model
was developed in subsection 3.5.3. Accordingly, one has a bonding and an
anti-bonding A1-level, and a bonding and an anti-bonding T2-level. From ab
initio calculations it is known that the bonding levels lie within the valence
band, the anti-bonding A1-level in the gap, and the anti-bonding T2-level in
the conduction band. Of the 10 electrons of the defect molecule, 8 can be
placed into the two bonding levels within the valence band. The remaining
two electrons just suffice to fill the anti-bonding A1-level in the gap. These
are the two electrons which, by exciting the crystal, are transferred to the
conduction band giving rise to the doubledonor nature of the center. How-
ever, the population of the two anti-bonding states is energetically costly.
Therefore it is not surprising that for the As atom other positions than the
substitutional one may be more favorable for minimization of the total en
ergy. In the above mentioned ab initio calculations, it was shown that a
displacement of the As atom by about I A in [Ill]-direction, a displacement
away of one of the 4 nearest neighbor A s atoms towards an interstitial po-
sition about 0.2 A below the plane spanned by the other three As atoms
(see Figure 3.22), wsiilts in a relative totdl energy minirrium which lies only
0.25 eV above the absolute minimum of the substitutional As-site. In ad-
dition to the stable substitutional state of the defect, there exists, t h e r e
fore, a metastable interstitial state. Thc stable state ia separated from the
metastable by an energy barrier of about 0.8 eV, and in the reverse direction
the barrier amounts to about 0.34 eV. If onr of the two donor electrons at
the substitutional center is optically excited, the barrier decreases siibstan-
tially and a thermal transition into the metastable state is possible. In this
state the center is not capable of capturing the electron whir11 was previously
optically excited into the conduction baud. This electron remains there re-
sulting in the experimentally observed persistent photoconductivity. Only
by heating the sample above toom tempwature can the As atom return to
Its stable substitutional site.
The main reason for the relatively small energy difference between the
substitutional and interstitial locations of the As atom iu that, between an
3.5. Deep levels 333

Direction

Figure 3.22: Geometry of the stable and metastable state of the As-antisitc defect
in GaAs. (After Dabrowski and SchrJler, 1992.)

interstitial As atom and Ohe three adjacent, A s atoms, st,rong sy2-bonds can
be formed similar to the bonds in the graphite structure of carbon and that
of g a y As. The non-bonding p-orbital of the A s atom, which lics rcla-
tively high enmgetically, does not increase t,he total energy of thc interstif,ial
s h t e because it remains largely unoccupied, the two remaining electrons are
hosted by the dangling bond of t,hc fourth remotc lying As atom.

DX-centers

A metastability similar to that of the C:aAs:AsG,, ailtisite defect is also ob-


served at ot,hcr point perturbations, among them at t,he so-called L)X-cent,ers
in GaAs and (Ga,Al)As mixed crystals (Lang, Logan, 1977). The micro-
scopic nature of these centers was a puzzle for a long time. It was only
clear that they were related to impurity atoms or the main groups l V and
VI of the periodic table which normally are incorporaled subst,itiitionally on
cation and anion sitcs, respectively, and form shallow donors. lTndcr cer-
tain conditions hydrostatic pressure or strong Ti-type doping in the case of
GaAs, and AlAs mole fractions larger than 22% in the case of (Al, Ga)As -
deep levels emerge from these shallow donors. Originally this transition was
attributed to the formation of a defect complex involving the donor atom
(a)and ti11 iinknown point, perturballon (X). Now, it is clear t,hat the D X -
center j u s t represents another state ol 1he shallow donor. In contrast to the
lat,t,er,the donor atom of a DX-ccntcr is not neutral, but singly negatively
charged, and its installation does not occur at a substitutional site but at an
interstitial site. If the special Conditions mentioned above are not fuliilled,
t,hen the shallow donor represents a stable state of the impurity atom, the
DX-state is metastbble. However, as for the EL2-center, thc rclalive total
energy minimum of the DX-state lies only slightly (about 0.2 e V ) above
thc absolute minimum of the shallow donor state. The reason for this en-
ergy balance is the sainc as in the E L 2 case, namely the energetically costly
populalion of the anti-bonding A1-level at the shallow donor. In the singly
negatively charged state which the donor takes under high doping, (and ev-
334 Chapter 3. Electronic structure of semiconductor crystals with perturbations

idently also under the other above-cited conditions), 2 electrons must be


hosted by this level. The energy costs thereby become so high that the DX-
state is energetically more favorable for the impurity atom than the shallow
donor state. The donor atom - to be concrete we will identify it below with
a Si atom - moves into the interstitial position shown in Figure 3.22, where
it can enter into sp2-bonding with the 3 surrounding As atoms. The non-
bonding porbital of the Si atom remains largely unoccupied, instead the
more deeply lying dangling hybrid orbital of the fourth more remote lying
As atom becomes occupied (Scherz, Scheffler, 1993).
This model explains in a natural way why the DX-center exhibits persis-
tent photoconductivity. By optically exciting an electron at the DX-center
into the conduction band, the center passes into its neutral charge state.
This means that the substitutional shallow donor is a stable state of the
point perturbation. Therefore, optical excitation transfers the DX-center
into a shallow donor in its ground state, and the donor is not able to cap-
ture the excited electron from the conduction band. It remains there for a
longer time and gives rise to persistent photoconducting. Only after thermal
excitation does the center return to its original metastable state.
The structural metastability observed at the DX- and EL2-centers is
not restricted to these point perturbations, but represents a relatively com-
mon phenomenon in tetrahedral semiconductors. It arises, evidently, from
the fact that (depending on the number of electrons to be placed at the
center), either the sp3- or the sp2-hybridization of the s- and p-orbitals of
the impurity atom allow for a lower total energy of the perturbation center,
jointly with the fact that the two kinds of hybridizations lead to different
atomic structures - the sp3-hybrids to the tetrahedral diamond structure,
and the sp2- orbitals to the graphite structure. Whether or not a metastable
state actually exists for a particular point perturbation can, however, only
be decided by ab initio calculations.

3.6 Clean semiconductor surfaces

3.6.1 The concept of clean surfaces


Every solid is bounded by a surface. Nonetheless, the model of an infinite
solid which neglects the presence of a surface works very well in many cases.
Why is this possible? The reason is, first, that in many cases one deals
with properties, such as transport, optical, magnetic, mechanical or thermal
properties, to which all the atoms of the solid contribute more or less to
the same extent, and, secondly, that there are many, many more atoms in
the bulk of a solid than at its surface, provided the solid is of macroscopic
size. In the case of a silicon cube of 1 cm x 1 c m x 1 cm, for example,
3.6. Clean semiconductor surfaces 335

one has 5 x loz2 bulk atoms and only 4 x 1015 surface atoms. Beside the
above mentioned bulk properties, there are other properties or processes,
like crystal growth, oxidation, etching or catalysis, which are determined by
the surface atoms only. In these rases, the model of an infinite solid does not
apply, of course. Moreover, in semiconductors, bulk properties like transport
are often not controlled by all atoms but only by dopant atoms. Then the
number N s of surface atoms per cni i s to be compared with the number
of dopant atoms. For a semiconductor sample of thickness d , area 1 cm2,
and doping concentration N D , the number N o x d of dopant atoms per c7n2
can easily become as small as the number IV, of surface atoms. This means
that the transport properties of such a semiconductor sample will depend
on its surface. In the history of semiconductor physics, this was recognized
at a very early stage. Examples which demonstrate this are the electric
rectification at a semiconductor-metal contact (discovered by F. Braun in
1874), and the unsuccessful attempts to build a field effect transistor in
the late thirties of this century. The failure was caused by electron states
localized at the surface which captured all the electrons induced by the
extcrnal voltage.
lhe surfaces used in the early field effect experiments were far from
perfect. They were made by cutting or cleaving a semiconductor sample in
air. If at all, they were clean and smooth in a macroscopic sense, but not
so microscopically. They exhibited surface roughness on the 100 nm scale,
structural defwts on the 1 nm scale, and impurity atoms at and below the
surface. The surface states responsible for the difficulties of the early field
effect transistor were due to these imperfections. In the present section we
will deal with surfaces which are free of such imperfections. Perfect surfaces
of semiconductor crystals in this sense necessarily represent a particular
lattice plane occupied only by chemically correct atoms at regular sites.
No impurity atoms are allowed above or below the surface. The surface is
reduced to what it means ideally, the termination of the crystal.
Perfwt surfaces in this sense cannot, of course, be realized in practice.
One may only try to approximate them so closely that the existing imper-
fections do not essentially change the properties of the surface as compared
to the properties which a perfect surface would have, if it really could be
made. One calls such almost perfect surfaces clean. Although this term
refers only to chemical composition, it also implies structural perfection or
atonzc smoothness. There are essentially t h r w ways to manufactwe clean
surfaces. All thrw need ultrahigh vacuum (UHV), i.e. pressures below
Torr:
(i) Treatment of imperfect surfaces by ion bombardment and thermal an-
nealing (generally in several cycles).
(ii) Cleavage under UHV conditions (only surfaces which are cleavagc planes
336 Chapter 3. Electronic structure of semiconductor crystals wilh perturbations

of the crystal can be made in this way, of course).


(iii) Epitaxial growth of crystal layers by means of molecular beam epitaxy
(MUE, see section 3.7).
Although most surfaces used in practice are not clean but imperfect, studies
of clean surfaces are also of practical importance. This is due to the fact
that imperfect surfaces are such complex subjects that it is very difficult
to approach them directly. One first has to reduce the level of COEApleXity
by considering clean surfaces for investigation. Later, the complexities are
introduced step by step and examined for changes. In this way, real imperfect
surfaces may also be undeistood. Clean surfaces are, howevei, important
themselves. It was already pointed out that these are the surfaces upon
which epitaxial crystal growth proceeds in MUE.
In this section, we will deal with the atomic and electronic structure
of clean semiconductor surfaces. The basic principles will be treated in
subsection 3.6.2 for atomic structure and in subsection 3.6.3 for electronic
structure. Particular suxfuces are discussed in subsection 3.6.4 taking Si and
GaAs as examples.

3.6.2 Atomic structure of clean surfaces


A geometrical construction which is of special significance in describing crys-
tal surfaces is that of lattice planes, which we will now describe.

Lattice planes of 3D crystals

Such planes are usually denoted by Miller indices ( h k l ) where h, k , 1 are


the integer reciprocal axis intervals given by the intersections of the lattice
planes with the three crystallographic axes. The symbol (loo), for example,
dcnotes lattice planes perpendicular to the cubic x-axis, (111)means lattice
plunes perpendicular to the space diagonal in the first octant of the cubic
unit cell, and (110) denotes lattice planes perpendicular to the face diagonal
in the first quadrant of the ry-planc of the cubic unit cell. In the case of
trigonal and hexagonal lattices, four crystallographic axes are considered
(three instead of two perpendicular to the c-axis). The lattice planes then
are characterized by four indices ( h k i l ) instead of three. The first three,
+ +
however, are not independent of each other, since i h k = 0. The (hkil)
are sometimes termed Bravais ~ T L ~ ~ C C R .
A particular geometrical plane can also be characterized by its normal
direction. To define lattice planes in this manner, it is convenient to write the
normal direction as a linear combination of the primitive vectors bl,b2, b3
of the reciprocal lattice introduced in scction 2.4, with integer coefficients
h l , Ik2, jL3,
3.6. Clean semiconductor surfaces 337

n - h l b l + h2bz +h ~ b 3 . (3.182)
Hrre, reciprocal lattice vectors are understood without the factor 27~in def-
inition (2.122). If the normal direction n is given, the roefkients ( h l h ~ h 3 )
are only determined up to a common integer factor. We chose this factor
such that h l , h 2 , hs have no common divisor. Then the coeffirients hi,h2, h 3
are the Miller indices of the lattice plane under consideration, however, r e
ferring to the primitive lattice vectors al,a 2 , a 3 as coordinate axes rather
than to the crystallographic axes. The latter are parallel to the piimitive
lattice vectors only in the case of primitive Bravais lattices. For centered
laltires, like the face rentrred cubic one, they have different directions, and
the coefficienls ( h l h ~ h 3differ
) from the corrirnon Miller indices. Tf necessary,
one can rasily switrh from one representation to the other. The ( h l h 2 h 3 )
are, apart from a co~linionfactor, obtained by multiplying the ( h k l ) with the
matrix which transforms the three nun-primitive crystallographic axis vec-
tors into thc three primitive lattice vectors. Using the above characterization
of lattice planes by their normal direction n, the lattice points

Ro = rloa1-k r20az + '7'30ag (3.183)


of the lattice plane perpendicular to n which contains the zero point, may
be defined by the equation

n & E h17.10 + h2r20 + h3r30 = 0. (3.184)


The unique character of this equation lies in the fact that only integer so-
lutions r10,r2O1r3o are allomwl. In mathematics it is called a Dzophantin
equation.
Whereas the Miller indices definr an infinite family of parallel lattice
planes, equation (3.184) yields only a single plane, namely that member of
the infinite family which contains the zero point. One can show that all
lattice planes

Ri = w a l + v z a 2 +w a ~ (3.185)
of the infinite family of parallel plancs are obtained by replacing the right
hand side of quation (3.181) by arbitrary integers 1,

(3.186)
The points of the lattice plane defined by equation (3.184) form a 2D lattice.
The primitive vectors of this lattice will be denoted by fi and f2. This i s
done in such a way that the three vectors f1, f2, n form a right hand coor-
dinate system. The f1, f2 may be expressed in terms of the primitive lattice
338 Chapter 3. Electronic structure of semiconductor cryst& with perturbations

Table 3.8: Primitive surface lattice \rectors and stacking vectors of low index sur-
faces of semiconductor crystals.

n Crpst'alStructures

I
I1 Diamond, Zincblende, Rocksalt.

1 WurtBite, Selenium

vectors a l , a z , q of the crystal under consideration. In order to determine


the coefficients of this representation, the Diophantin equation (3.184) has
to he s o l v d , which ran be done by Incans of the Euclidean algorithm (set,
Bechstrdt and Enderlein, 1988j. The piirriitive vectors fi, fz obtained in this
way are shown in Table [3.8) for several Iow index lattice planes of the 5
common scrniconductor crystal structures. IJsing thc lattice vectors fi. f2
and ailitrary integers ~ 1 ~ 5 the2 , lattice plane given by relations (3.183) and
(3.184) can bc represented as

Ro = S l f i + sgh. (3.187)
The lattice planes Rr defined by equations (3.185 and (3.186) can be written
in the form

where f.3 is a vcctor complementing F1 and f 2 to form a s ~ of


t primitive lattice
vrctors Fi, f2, F3 of the 313 bulk lattice of the crystal. The vector 3 can br
detcrinined from thr Diophantin equation (3.186) for 1 1 in just the same
~

way, i . ~ .by means of the Euclidean algorithm. BS the vectors Fi,fid were
obtained above. The choice of $3 is not uniquc. of course, and any vector fi
which differs from f 3 hy a vector aithiii the latticr plan? can also be used.
We call f 3 the stacking vector because it determines how the lattice planes
arc stacked in the crystal.
The considerations above show that the primitive unit cell of a crystal
may be chasm as a parallelrpiyed with one of its pair6 of parallel laws par-
allel to a gwen lattice plane. This implies that the structure of a crystal
3.6. Clem semiconductor surfaces 339

Figure 3.23; Construrtion of a crystal from its lattice planes. The lattice points of
a given plane are occupicd by identical atoms. Different planes may host different
atoms if the crystal has a basis. In the figure, a crystal with o~ilyone atom per
primitive unit cell is shown.

may be characterized as consisting or parallel lattice planes which are dis-


placed with respect to each other and which are each o c c i i p i d hy atoms
of a particular species (see Figure 3.23). 'I'o be more specific, the lattice
plane Ro is occupied by atoms of species 1, the next plane, displaced by 6
with respect to the first onc, is occupied by atoms of species 2, etc., and the
plane displaced by FJ is occupied by atoms of type J . It may happen that
two OT more atoms of the basis are located at the same lattice plane. In
that case an ntomic layer consists of two or more basis atoms. The lattice
+
plane Ro 7j completev the constsuction of a crystal slab which, in the
vertical direction, encompasses exactly one yrimitivc bulk unit cell. This
slab is callcd a p r i m i t i v e crystal slab. A lattice plane occupied by atoms is
referred to nb: a t o m i c layer. The second primitive crystal slab begins with
a lattice plane occupied by atoms of species 1 and is displaced with respect
to the zero-th plane of the first layer by the stacking vector f3, followed by
a plane with atoms of species 2 which is displaced by f3 t ?2 etc., the last
plane of the second primitive crystal slab being occupied by atoms of species
+
J and displaced by f3 T'J, The second primitive crystal slab is followed
by another J planes dlvplaced with respect to the first slab by 2f.j instead
of f3. The crystal can therefore be thought of as consisting of successive
primitive crystal slabs sitiihtcd one above the other, each of which consists
of atomic layers containing difkrmt types of atoms and which are laterally
displaced with respect to onc another. The location of an individual atom
can bc specified by the number 1 of the primitive crystal slab, the numbers j
of the atomic sublattice and the integer coordinates s1, s2 witahinthc lattice
plane. The position R(j, I , s1, s2) of an atom j in crystal slab 1 can therefore
be written as
340 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.189)
The complete set of locations of atoms in an infinite 3D crystal can be
obtained by assigning all possible integer values from --oo to +m for 1, s 1 , s 2 ,
and all integer values from 1 to J for j . The normal direction n, upon which
the whole construction of lattice planes is based, is arbitrary in the case of
an infinite 3D crystal. Any choice of n yields the same crystal.
We can immediately employ the above representation of an infinite crys-
tal in describing a crystal with a surface. The normal direction n is then,
however, fixed by the direction perpendicular to the surface. We start with
the description of a crystal having an ideal surface. The exact meaning of
ideal in this context is explained immediately below.

Ideal crystal surfaces

Regular crystal sites

We consider a crystal surface given by a lattice plane perpendicular to n


as defined by equation (3.184). A crystal having such a surface may be
generated from an infinite crystal of infinite extension by removing all atomic
layers above the surface and retaining those below. Since the forces acting on
atoms situated beneath a lattice plane in an infinite crystal are also partially
due to the atoms located above the plane, we can, in general, expect that the
forces acting on atoms in a crystal with a surface should differ from those
acting in an infinite crystal. The deviation from the infinite case, however,
diminishes with increasing distance of atoms from the surface and we can
thus assume that the forces acting on, and hence the positions of, atoms
deep inside the crystal bulk are, to a good approximation, the same as those
in an infinite crystal. This is, however, not true for atoms situated near the
surface, and the forces acting on them are appreciably different, resulting in
displacements of atomic positions with respect to those of the infinite crystal.
We will discuss these displacements below. Here, we assume that they are
not present and, correspondingly, the atoms at and immediately below the
surface have the same positions as they would in an infinite crystal. The term
ideal surface is used to refer to this configuration. The atoms of a crystal
having an ideal surface are thus located at the positions R(j, 1, ~ 1 , s ~given
)
by equation (3.189), however, only positions below the surface plane are
occupied. These obey the relation

n . R(j, 1 , s 1 , 5 2 ) = I + n . Tj 5 0. (3.190)
The surface or first atomic layer is obtained if the left hand side of this
relation is taken to be zero. A solution of (3.190) is 1 = 0 and = 0.
3.6. Clean semiconductor surfaces 34 1

The latter condition signifies j = 1 since 71 = 0 is assumed. Thus, the


first atomic layer corresponds to the particular lattice plane perpendicular
to n which goes through zero and whose lattice points are occupied by basis
atoms of species 1. There may be other ?j beside ?I which, although not
being zero themselves, have a zero projection n .3with respect to n. Then
the basis atoms of this species j are also located in the first atomic layer.
They are shifted with respect to the atoms of species 1 by a vector Tj parallel
to the surface. Such multiple-species occupancy of an atomic layer occurs,
for instance, in the case of (110) surfaces of diamond and zincblende type
crystals. In this case one has two atoms in each primitive unit cell of the 2D
lattice of a lattice plane, in the case of Si, for example, Si atoms, and in the
case of GaAs, 1 Ga atom and 1 As atom.
The distinguished role of atomic species 1 in the above considerations
results from the choice of the origin - it has been placed at an atomic site
of species 1. Of course, the coordinate system may be shifted in such a way
that its origin coincides with the location of any other basis atom. In each
case a different surface is obtained. Even if the basis atoms are chemically
identical the surfaces may differ from each other in a topological sense. In
the case of diamond type crystals, for instance, two topologically different
(111) surfaces exist, one with three nearest neighbors above and one below
the surface and another with one above and three below. The latter surface
is more stable than the former one and the latter is meant if one refers
to a (111) surface. For other crystal structures and surfaces the situation
is similar. If topologically different surfaces exist for a given set of Miller
indices, generally, one of them is more stable than any other, and that is the
one which is commonly realized in experiment and studied theoretically.
We first consider the translation symmetry of an ideal surface and the cor-
responding lattice.

Translation symmetry and lattices of ideal crystal surfaces

The translation symmetry of a crystal with an ideal surface can be derived


from the following observations:
(i) Only translations within the surface lattice plane perpendicular to n are
admissible. Any translation leading out of this plane would alter the spatial
location of the surface and thus would not transform the system crystal
with surface into itself. The symmetry group of translations and, corre-
spondingly, also the lattice of a crystal possessing a surface are, therefore,
only 2-dimensional, although the crystal with surface is 3-dimensional in
extent.
(ii) Only those translations are admissible which transform each atomic layer
into itself. The construction of a crystal, layer by layer, as described above
implies that if a particular translation transforms the first atomic layer into
342 Chapter 3. Electronic structure of semiconductor crystals with perturbations

itself, this also holds for any other layer. It immediately follows that the
group of translations of a crystal with a surface is identical to the translation
group of the first atomic layer. The lattice vectors r of a crystal possessing
an ideal surface perpendicular t o the normal direction n are therefore those
of equation (3.187), so that

r =slf1+ s2f2 (3.191)

holds. The vectors f1, f 2 are the primitive lattice vectors of the 2D lattice
of the ideal surface. The primitive unit cell of the crystal with surface has
the form of a prism bounded above by the parallelogram spanned by f1, f 2
and extending to minus infinity in the direction of -f3. For the low-index
surfaces of the common semiconductor crystal structures, the three vectors
f1, f2, f 3 are listed in Table 3.8.
As in the 3D case, 2D lattices may be divided into crystal systems and
Bravais lattices according to their symmetry with respect to rotations and
reflections. We now consider the possible plane crystal systems and plane
Bravais lattices. Their point symmetry elements are necessarily rotations
about axes which are perpendicular to the 2D lattice plane, and reflections
at lines within the 2D lattice plane. Thus, the point groups are either C ,
( n ) or C,, (nm, nmm). Since the rotation through 180' is always a sym-
metry element of a plane lattice, only even n are allowed. The highest value
of n may be readily obtained from the derivation of the possible rotation
symmetry axes of 3D crystals in Chapter 1. There, n 5 6 was found. Thus,
the possible multiplicities of rotation symmetry axis of plane lattices are
n = 2,4, and 6.
A lattice which only contains a 2-fold symmetry axis is either a com-
pletely general oblique lattice or a rectangular lattice. The point groups of
these lattices are, respectively, C 2 (2) and C2, (2mm). Lattices with a 4-fold
symmetry axis also possess 4 reflection lines which are rotated through 45'
with respect to each other. The point group of such a lattice is therefore C4v
(4mm). Similarly, lattices with a 6-fold axis have 6 reflection lines which
meet at an angle of 30'. In this case the point group is c 6 v (6mm). There
are thus 4 different plane crystal systems - the oblique with holohedral point
group 2, the rectangular with holohedral point group 2mm, the quadratic
with holohedral point group 4mm and the hexagonal with holohedral point
group 6mm (see Figure 3.24).
The possible plane Bravais lattices are obtained as follows. First, one
takes the four primitive lattices with unit cells which are parallelograms
having either no particular symmetry, or that of a rectangle, a square or an
equilateral hexagon. Then one adds additional points to each of the unit cells
of these lattices in such a way that their point symmetries are not lowered.
Only one new lattice is obtained in this way, namely the 'body' centered
3.6. Clean semiconductor surfaces 343

oblique rectangular quadratic hexagonal


P P P P

Figure 3.24: The 5 plane Bravais lattices of the 4 plane crystal systems.

rectangular lattice. This lattice cannot be transformed by continuous and


symmetry preserving transformations in the primitive rectangular lattice,
nor in any other primitive lattice. So it forms an additional Bravais lattice.
In all other cases, the addition of points either leads back to the primitive
lattice or results in no lattice at all (i.e. it creates a crystal with a basis). We
conclude that five plane Bravais lattices can be realized within the framework
of the 4 plane crystal systems: only the primitive lattice in the oblique case,
the primitive and the centered in the rectangular case, and again only the
primitive in the quadratic and hexagonal cases (see Figure 3.24).
With the last statements we have completed the symmetry classification
of the plane lattices of crystal surfaces. The lattice types of the ideal low
index surfaces of the common semiconductor structures are summarized in
Table 3.9. We now turn our attention to the symmetries of crystals with
surface as a whole.

Point and space group symmetries of ideal crystal surfaces

We wish to establish, on the one hand, point groups which transfer equivalent
directions of a crystal with surface into one another and, on the other hand,
space groups which transform the crystal with a surface into itself. In the
latter case this implies that not only does the 2D lattice transform into
itself, but so do equivalent atoms occupying the primitive unit cells. These
atoms are located at positions R(j,I, sir92) given by equation (3.189). To
express the 2D nature of the translation symmetry of a crystal with a surface
explicitly, we denote the atomic positions by Rjl(s1, s2) R(j,I , ~ 1 , s and
~)
write
344 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Table 3.9: Structural properties of ideal low index surfaces of the five common
semiconductor crystal structures. Column 6 and 7 give, respectively, the number
of atomic layers of an irreducible crystal slab and the number of basis atoms per
layer.

Bravais Lattice Point Group Space Group Irr. Slab Basis

hexagonal 3m p3ml 6 1

square 2mm p2mm 4 I

p-rectangular 2mm p2mg 2 2


hexagonal 3m p3m1 6 1

square 2mm p2mm 4 I


p-rectangular m plml 2 2

hexagonal 3m p3ml 6 1

square 4mm P h m 2 2
prectangular 2mm p2mm 2 2
hexagonal 3m p3ml 4 1

(ioio) prectangular m plml 4 2

prectangular m P W 4 4

hexagonal 1 Pl 3 1

(ioio) p-rectangular 1 Pl 6 1

(1120) p-rectangular 2 P21 4 1

where

represents the basis of the crystal with surface. This basis contains an in-
finite number of vectors corresponding to the infinite number of atoms in a
primitive unit cell. The point and space group elements of a crystal with
surface must leave the surface invariant, i.e. only 2D symmetry groups need
to be taken into consideration. Here, as in the case of translation symmetry,
we therefore also have the situation that, although a crystal bounded by
a surface extends in three dimensions, its point and space groups are only
2-dimensional.
The point groups of directions of crystals with plane surfaces are nec-
essarily subgroups of the holohedral point groups of the plane lattices, i.e.
subgroups of the point groups 2,2mm, 4mm and 6mm. There are exactly
3.6. Clean semiconductor surfaces 345

a @
..
@ ..

0c3
2

1
2 mm

m
(Ii
.4 m m

4
a 6mm

@ 3m

(I>
3

Figure 3.25: The 10 point groups of equivalent directions of crystals with plane
surface.

10 such groups. Their stereograms are shown in Figure 3.25. This implies
the existence of 10 different crystal classes which are associated with corre-
sponding crystal systems as indicated in Figure 3.25.
The possible plane space groups can be found as follows (see Figure 3.26).
To start with, it is evident that each of the 10 point groups of equivalent
directions combined with the corresponding associated lattice gives rise to a
space group. The space groups p l , p 2 1 1 , p l m l , p 2 m m , p 4 , p 4 m m , p 3 , p 3 m l , p 6
and p 6 m m originate in this manner. Since the point groups of the rectan-
gular crystal system are each associated with two Bravais lattices, primitive
and centered, we find two further space groups, c l m l and c 2 m m . In the case
of the point group 3 m , there are two different possibilities of positioning the
reflection lines relative to the hexagonal lattice vectors, either through the
vertices of the equilateral hexagon of the Wigner-Seitz cell, as assumed in
the case of p 3 m 1 , or such that ,hey bisect its edges. In the latter case one
346 Chapter 3. Electronic structure of semiconductor crystals with perturbations

-
crystal
svstem
Bravais
lattice
ooint soace 1

oblique
2
LLZ

rec- p-rec-
t angular
2mn
t a n g ul o r
6
I I
p'g'

Ip- I
c-rec-
tangular

Figure 3.26: The 17 space groups of crystals with plane surface.

has the group p31m as a 13-th space group. One should further note that
the point group of directions of a crystal remains unchanged if, in its space
group, a glide reflection line is substituted for an ordinary reflection line. One
must therefore examine the 13 space groups already established to determine
whether the substitution of a reflection line m by a glide reflection line g (i.e.
a reflection in m in conjunction with a translation 7' by half of the shortest
lattice vector parallel to m ) leads to a new space group. One easily finds
that this is not the case for the hexagonal crystal system. In the quadratic
crystal system it is possible to substitute a system of glide reflection line for
one but not both of the non-equivalent reflection line system. This yields the
additional space group p4mg. The additional space groups in the case of the
3.6. Clean semiconductor surfaces 347

2mn

c2mn I
I
p-square 4

- p4

4mn

p4mn I
-
,-hexa- 3
gonal

p3m1 I

Figure 3.26: Continuation: The 17 space groups of crystals with plane surface.

primitive rectangular crystal system are p l g l (from p l r n l ) and p2rng,p2gg


(from p2mrn). The centered rectangular and the oblique crystal systems do
not give rise to additional space groups. In the case of crystals with plane
surface, there are therefore a total of 17 different space groups. Four or them
involve glide reflections. i.e. they are non-symmorphic.
The 2D point and space groups of the various surfaces of a given crystal
may be derived from the 3D point and space groups of the infinite crys-
tal under consideration. They are, in fact, the subgroups of the 3D groups
which contain only those symmetry elements which transform lattice planes
348 Chapter 3. Electronic structure of semiconductor crystak with perturbations

situated parallel to the surface into themselves. More precisely, they con-
tain elements which transform directions in a lattice plane into equivalent
directions in that plane as far as point groups are concerned, and transform
atomic layers located parallel to the surface into themselves if space groups
are under consideration. The 2D point and space groups of the three low-
est index surfaces of the five common crystal structures are summarized in
Table 3.9, columns 4 and 5. The point and space group symmetry of an
ideal crystal surface may also be derived from the projection of the crystal
onto its surface. In Figure 3.27 such projections are shown for the three
low index surfaces of diamond and zincblende type crystals. Atoms from
different layers below the surface are illustrated differently with sizes as in-
dicated. After some number of atomic layers (6 for (lll),4 for (loo), and
2 doubly occupied layers for ( l l o) ) , the projections repeat themselves. A
crystal slab which contains the minimum number of atomic layers necessary
for completing the projection, is called an irreducible crystal slab. The point
and space group symmetries of a crystal with an ideal surface are those of
its irreducible crystal slab.

Relaxed and reconstructed surfaces

Surface-induced atomic displacements

In the preceding section the atomic structure of crystal surfaces was con-
sidered under the assumption that the atoms of the crystal bounded by a
surface occupy the same positions Rjl(s1,s2) as they did in the infinite 3D
crystal, if the former is generated from the latter by removing one half of
it. As already noted, this assumption is actually not valid. The atoms of
the surface layer experience different forces than those acting in the bulk
of the crystal, and are thus subjected to displacements from their original
sites in the crystal bulk. Since the forces acting on the atoms of the second
layer are in part determined by the positions of the atoms in the first, these
forces are also subjected to changes accompanied by displacements in the
second layer and so on for each successive layer. All one can assume is that
the displacements decrease from one layer t o the next and vanish altogether
at a depth that is relatively far from the surface. Here, we present a more
detailed description of the surfaceinduced displacements of atoms and dis-
cuss the resulting altered symmetries as compared to those of ideal crystal
surfaces.

Translation symmetry of relaxed and reconstructed surfaces

We denote the displacements of atoms due to the formation of the surface


by 6Rjl(sl, 4, and the new positions of atoms in the crystal with surface
3.6. Clean semiconductor surfaces 349

(1111

(1101 coiii

(111) (100) (110)

0 1 A 0 1 A 0 1 A
O2B 0 2 6 016
0 3A 0 3A 2A

Figure 3.27: Projections of a diamond or zincblende type crystal onto its low index
surfaces.

by Rj~(s1,
sz). Then

6Rjl(sl, s2) 4 0 f o r I -+ -m. (3.195)

The displacements 6Rjl may be divided into two classes with regard to their
effect on translation symmetry. If the latter is not affected, the displacements
are termed r-e-eluzation. In this case equivalent atoms in different primitive
unit cells are displaced in the same way, i.e.
350 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.28: Surface relaxation (a) and surface reconstruction (b) extending up to
the second atomic layer. A 2 x 2 reconstruction is shown in part (b).

BRjl(s1, s2) = 6Tjijl for all 51, s2. (3.196)


Only the vectors 61 of the basis are altered, the lattice vectors remain un-
changed in this case (see Figure 3.28a). If, on the other hand, the translation
symmetry is altered, the displacements are termed reconstruction. In this
case equivalent atoms in different unit cells are not all displaced in the same
manner, i.e. bRjl(s1,s2) depends on sl,s2. Both the basis and the lattice
are changed. We first consider the changed lattice translation symmetry of
reconstructed surfaces (see Figure 3.28b).
In describing reconstruction it is useful to divide the crystal with a surface
into two slabs parallel to the surface, an upper slab containing the atomic
layers with displaced atoms, and a lower slab encompassing all other layers,
i.e. layers with non-displaced atoms. The upper slab is sometimes called
a selvedge, here we use the terms surface slab and bulk slab (or simply
bulk) for the upper and lower slabs, respectively. By Ts and Tb we denote,
respectively, the plane translation symmetry groups of the surface and bulk
slabs. The translations which transform the crystal with surface into itself
must belong to both groups of translations, Ts as well as Tb. The translation
group T of the whole crystal with surface is thus the intersection

T = Ts fl Tb (3.197)
of Ts and Tb. Alternatively, one can say that T is the largest common
subgroup of both groups Ts and Tb. There are two possibilities, either T
only consists of the identity translation, which means that the lattices defined
by Ts and Tb are non-commensurate and the crystal with surface does not
possess any lattice translation symmetry, or T contains more elements than
just the identity, in which case one says that the two lattices derived from
Ts and Tb are commensurate. The lattice associated with the T is called the
coincidence lattice. If, in particular, Tsis a subgroup of Tb then T is equal
3.6. Clean semiconductor surfaces 35 1

to T,, i.e. the coincidence lattice is identical to the lattice T, of the surface
slab. If T, is not a subgroup of Tb then T cannot be equal to T, and is
necessarily a proper subgroup of T,, i.e. it is smaller than T,. One can thus
distinguish between the following three cases with regard to the translation
symmetry of crystals having a reconstructed surface:
(i) no translation symmetry,
(ii) translation symmetry exists but is smaller than that of the surface slab,
(iii) translation symmetry exists and is the same as that of the surface slab.
If one assumes that, among the various conceivable surface reconstructions
with a given degree of translation symmetry, that particular reconstruction
will take place which allows for maximum translation symmetry of the crystal
with surface (the system surface slab plus bulk slab), then only the third
of the above possibilities can be realized. A formal proof of this assumption
does not exist, and it is probably not valid without exceptions, however, as
a rule, it generally yields correct results.
Using the above conclusions we are able to determine the primitive lattice
vectors of the reconstructed surface in terms of the primitive lattice vectors
f1, f 2 of the ideal surface. The latter are, by definition, also the primitive
lattice vectors of the bulk slab of the crystal with surface. Let be fi and fi
the primitive lattice vectors of the reconstructed surface slab. They may be
linearly composed of f 1 , f 2 according to

f: = Qllfl + 912f2 (3.198)

f; = Q2lfl + 922f2. (3.199)


Here, the coefficients Q i k , i, k = 1,2, of the transformation matrix Q are, at
the outset, arbitrary real numbers. They need to be rational if and only if
the two lattices derived from f i , f 2 and f l , f2 have a coincidence lattice, i.e.
in case (ii) above. If the coincidence lattice is identical to the lattice derived
from fi, fi, i.e. in the particularly interesting case (iii), it follows that the
Qik must have integer values. In this case, the lattice derived from f : , f i
is simultaneously the lattice of the reconstructed surface. We thus arrive
at the important conclusion that in the only case of practical interest (iii)
above, surface reconstruction can be described by a 2 x 2 matrix with integer
elements.
The two most common forms of this type of reconstruction have a special
notation (Wood notation):
(1) The non-diagonal elements vanish, i.e. f: and fi are parallel to f 1 , f 2 ,
respectively, and their lengths are integer multiples of the respective lengths
of the latter. We thus have
352 Chapter 3. Electronic structure of semiconductor crystak with perturbations

f: = nfi, f& = mf2 , (3.200)


with n and m being integers. The primitive unit cell of the surface slab con-
tains n x m primitive unit cells of the bulk slab. This is said to constitute an
n x m reconstruction An n x m reconstructed crystal surface of a particular
material C parallel to a lattice plane with Miller indices (hlcl) (or ( h k i l ) in
the case of hexagonal symmetry) is characterized by the symbolic notation

C(hlc1) n x m. (3.201)

(2) The off-diagonal elements of Q are not equal zero, i.e. fi is not parallel
to fi and/or fi is not parallel to f2. The angles L(f!, fi) and L(f&fz), which
are in general not equal to each other, are assumed to be equal in the case
under consideration. This means that the two vectors f1, f2 (with tails joined
at the same point) can be transformed into the two corresponding vectors
f;, fi by a rotation through the same angle L(f;, fi) = L(f4, f2) = a about
an axis which is perpendicular to the surface, with a subsequent rescaling
/ ~ f ~ ~ A, (hlcl) surface
of f1, f2 by the factors lfiI/fll and ~ f ~ ~respectively.
of a particular material C reconstructed in this way is characterized by the
symbolic notation

lfil x
C(hkl) - 141 - a.
- (3.202)
lfil If21
The factors lf~l/lf~land lfil/lf21 are in general irrational in contrast to the
qik, which in the case considered here are integers. Examples of both of the
special reconstruction forms discussed, as well as for the general reconstruc-
tion form in case (iii), are shown in Figure 3.29.
Sometimes, in the notation (3.201), n x m is replaced by p - n x m
or c - n x m. The lattice vectors fi = nfi and fi = mfz are then not
necessarily primitive as originally assumed in the notation (3.201), and in
addition to primitive ( p ) reconstructed surface lattices, also centered ( c )
ones are possible. This can only take place, however, for rectangular surface
lattices. Thus, the modified notation applies only to this case, although
it is also sometimes used (formally incorrectly) for square lattices. In the
rectangular case the notation c - n x m describes a type of reconstruction
which is not covered by one of the two notations (3.201), (3.202), and which
can otherwise only be characterized by the 2 x 2 matrix Q itself. For square
reconstructed lattices the c - n x m notation is just a simpler description of
a reconstruction of type (3.202).
The point symmetry of a reconstructed surface lattice is generally lower
than that of the ideal surface lattice from which it is derived. This im-
plies that the crystal with reconstructed surface belongs to a different plane
3.6. Clean semiconductor surfaces 353

Figure 3.29: Three different types of surface reconstructions: (a) 1 x 2, (b) 3h x


31 -30, (c) general type, the Wood notation does not apply in this case, the matrix
notation does with 411 = 5, 412 = -1, 421 = 2, 422 = 2.

crystal system than the crystal with ideal surface. For example, the 2 x 1
reconstruction of a square lattice leads to a rectangular lattice. The same
holds for the 2 x 1 reconstruction of a hexagonal lattice which also results
in a rectangular lattice.

A further item is worth mentioning, concerning the surface reconstruc-


tion itself. It follows from the point symmetry of the crystal with ideal
surface itself. If the latter belongs to the square crystal system, i.e. if it
has a square lattice and one of the two point symmetry groups 4 m m or 4,
the directions of the two primitive lattice vectors are symmetrically equiv-
alent. A particular reconstruction which increases the surface unit cell in
the direction of fl by a factor n and in the direction of f 2 by a factor m , is
equivalent to another reconstruction which does the same for, respectively,
the symmetrically equivalent vectors f2 and fi (Figure 3.30). An analogous
statement holds for an ideal surface of the hexagonal crystal system, having
a hexagonal lattice and one of the point groups 6mm, 6 , 3 m or 3. In this
case, three symmetrically equivalent direction exist (Figure 3.30). If there is
no physical reason which makes one of the different symmetrically equivalent
reconstructions more likely than another, they will take place simultaneously
in different regions of the surface. The result is the formation of domains
of otherwise identical, but differently oriented, reconstructed unit cells. Due
to the domain structure, the overall translation symmetry of the surface is
destroyed. Structural imperfections of a more local nature occur where the
boundaries of such domains meet.
354 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.30: Symmetrically equivalent 2 x 1 reconstructions in the case of ideal


surfaces belonging to the square (left) and hexagonal crystal systems.

Point and space symmetries

The point and space symmetries of relaxed or reconstructed surfaces are gen-
erally of lower degree than those of the corresponding ideal surfaces. They
do not only depend on the point and space symmetries of the irreducible
crystal slab as in the case of an ideal surface, but also on the point and
space symmetries of the relaxed or reconstructed surface slab. If the surface
slab consists of more layers than the irreducible crystal slab, then its point
and space groups can be taken to determine the point and space groups of
the whole crystal with relaxed or reconstructed surface. If the surface slab
contains fewer layers than the irreducible crystal slab it is expedient to add
further atomic layers (which then do not contain displaced atoms) to make
up the difference. The point and space group elements of the thus extended
surface slab which are simultaneously point and space group elements of
the irreducible crystal slab form, respectively, the point and space groups of
the whole crystal with relaxed or reconstructed surface. Similarly like this
was done for the translation symmetry group above, one may argue that the
point and space groups of the relaxed or reconstructed surface slab should be
subgroups of the point and space groups of the corresponding ideal surface.
If this is the case, the point and space groups of the relaxed or reconstructed
surface slab are, respectively, the point and space groups of whole crystal
with relaxed or reconstructed surface.

3.6.3 Electronic structure of crystals with a surface

The electrons and cores of a crystal with a surface undergo the same interac-
tions as the electrons and cores of an infinite bulk crystal, the only difference
being that the cores and electrons of the removed semi-infinite crystal above
the surface are missing, and the positions of the cores just below the surface
3.6. Clean semiconductor surfaces 355

differ from those in the infinite bulk. The two basic approximations of the
theory of the interacting electron-core system of an infinite crystal, namely
the adiabatic approximation and the one-electron approximation, are not
influenced by these modifications, so they may also be used in the presence
of a surface.
In the electronic structure calculations of infinite bulk semiconductor
crystals, the core positions commonly are taken as input data. This is pos-
sible because these positions are crystal sites of high symmetry which are
well-known from X-ray diffraction experiments. For crystals with a surface,
this can no longer be assumed. The positions of atoms in the surface layers
of relaxed or reconstructed surfaces are crystal sites of lower symmetry. In
many cases, they are not known or only incompletely known from experimen-
tal investigations, for these investigations are more difficult and less precise
than X-ray diffraction in the case of bulk crystals (a short review of these
methods will be given further below). X-ray diffraction cannot be applied
to surfaces because it lacks surface sensitivity. In such circumstances, the
positions of atomic cores at surfaces are to be treated as output, rather than
input, data of the electronic structure calculations. The way that this can be
accomplished was discussed in section 2.2 in general terms, and in section
3.5 with respect to point defects. It involves, first, the calculation of the
total energy of the electron-core system for a variety of different sets of core
positions and, second, the minimization of the total energy with respect to
these sets. The minimum set gives the core positions which really apply. The
most critical part of the total energy is the energy of the electron system.
In order to obtain it, the one-electron energies of the crystal with surface
have to be calculated for assumed core positions. Formally, this involves the
same task as in the case of bulk crystals, namely, the calculation of station-
ary one-electron states for given core positions. Below, we will demonstrate
how this problem can be solved in the case of crystals with surfaces. The
remaining parts of the procedure for determining the atomic and electronic
structures of surfaces, the calculation of the entire total energy including the
core-core interaction energy, and the minimization of the total energy with
respect to the core positions, will not be treated here because it is mainly a
numerical task.
The assumption of a priori known core positions is valid if ideal surfaces
are considered. In this case they are the infinite bulk positions. The elec-
tronic structures of ideal surfaces are important as reference data for the
electronic structures of relaxed and reconstructed surfaces. Thus we will
also deal with them.

One-electron Schrodinger equation

The one-particle Schrodinger equation for the wavefunction $E(x) of an


356 Chapter 3. Electronic structure of semiconductor crystals with perturbations

electron in a crystal with surface has the same general form

(3.203)

as in the case of an infinite bulk crystal (see equation 2.53). The oneelectron
potential V(x) may be written as the sum

of the electron-core interaction potential Vc(x), and an effective one-electron


potential Ve(x)due to electron-electron interaction. For Ve(x),any of the
one-electron approximations introduced in section 2.1 may be used. The
core potential Vc(x)follows from the corresponding expression for an infinite
crystal if the summation is restricted to cores within and below the surface.
+ +
Using the surface adapted notation 5 If3 r for the position of the j-th
+
basis atom of the primitive bulk unit cell at the bulk lattice point r If3,
j = 1 , 2 , . . . , J , 1 = 0, -1,. . . , -00, this leads to

-W J
(3.205)

where q(x) is the potential of a core of species j located at R = 0. The


core potential Vc(x)and, hence, also Ve(x)and V ( x ) have the translation
symmetry of the 2D surface lattice, so that

v(x)= V(x + r) . (3.206)


As in the 3D bulk case, this symmetry can be used to derive the Bloch
theorem.

Bloch theorem

This theorem states that the energy eigenfunctions $E(x) of the Schrodinger
equation (3.203) may be chosen simultaneously as eigenfunctions of the sur-
face lattice translation operators t,.. This allows one to write these functions
in the form of Bloch functions @Q(x) with a 2D quasi-wavevector

4 = 91g1 + 9282, (3.207)


where q1,92 are arbitrary real numbers, and gl and gz are the primitive
lattice vectors of the reciprocal surface lattice, defined by the relations
3.6. Clean semiconductor surfaces 35 7

fi . g k = 27T6&, i, k = 1,2, (3.208)


The latter equations are solved by the vectors

g1 = N-lf2 x [fl x f2], g2 = N-lfl x [f2 x fl] (3.209)

with N = (1/2n)[(f1 . fl)(fZ.fz) - (fi . f2)] as normalization constant. The


two vectors gl, gz of (3.209) lie in the plane spanned by the two primitive
surface lattice vectors f1, fz, i.e., within the surface. The same statement
applies to the 2D wavevector q.
As in the 3D bulk case it is convenient to introduce a region of macro-
scopic size with respect to which the eigenfunctions $J,E~(x) can be assumed
to be periodic. In the case of a crystal with surface this region forms a paral-
lelogram spanned by the edge vectors Gfl, Gf2, with G being a large integer.
The area RII of the periodicity region is G21f1 x f2l. The Bloch functions
normalized to it may be written as
1
Q,Eq(x)= -e2qx UEq(X), (3.210)
fi
where U Q ( X ) is the Bloch factor, which has the periodicity U E ~ ( X )= U E ~ ( X +
r) of the surface lattice and is normalized with respect to a primitive surface
unit cell. To guarantee the periodicity of the Bloch functions $ ~ ~ ( xof)
(3.210) with respect to the periodicity parallelogram, the wavevectors q must
have the form

(3.211)
with p1,pz as integers. This means that the q-vectors must belong to a
finely-meshed lattice similar to that of Figure 2.4. For the macroscopically
large values of G which we assume, q is practically continuous, although the
number of different q-values within a given region of q-space is finite.

Surface Brillouin zones and surface energy bands

The energy eigenvalues E of a particular Bloch type eigenfunction of equa-


tion (3.203) depend on q. If g varies over the whole infinite space, as we
assume here, E represents a unique function E ( q ) of q. This description
corresponds to the extended zone scheme in the case of an infinite bulk crys-
tal considered in section 2.4. As in the latter case, one may switch from the
extended to the reduced zone scheme in which q varies only over a primitive
unit cell of the reciprocal surface lattice. Any other point of the i n h i t e q-
space may be written as q + g where g is a surface reciprocal lattice vector.
358 Chapter 3. Electronic structure of semiconductor crystals with perturbations

In the reduced scheme, the energy function E(q+ g) over the whole q-space
is replaced by the manifold of functions E,(q) +
E ( q g ) defined only over
a primitive unit cell. As in the bulk case, there is a distinguished choice of
the primitive unit cell, as we will see below.
Using first order perturbation theory with respect to the periodic poten-
tial V(x), one may demonstrate that, in the extended scheme, E(q) repre-
sents a continuous function of q everywhere, except for the q-vectors with
lql = lq + gl or

q . g + -1 g2 = 0, (3.212)
2
where g is an arbitrary surface reciprocal lattice vector. Equation (3.212)
defines lines in the 2D q-space which are analogous to Bragg reflection planes
in the 3D k-space of a bulk crystal. On Bragg reflection lines, the energy
function E(q) has discontinuities. These lines may be used to define 2D
Brillouin zones in just the same way as was done in the 3D case in section
2.4. One speaks of surface Brillouin zones (BZs). The surface B Z s of the
square lattice have, in fact, already been used in section 2.4 as an illustration
for the 3D case. Of particular importance is the f i r s t surface B Z . It may
also be defined as the Wigner-Seitz cell of the reciprocal surface lattice.
Since there are 5 different plane Bravais lattices and, hence, 5 different
reciprocal surface lattices, there are also 5 different first surface B Z s . They
are shown in Figure 3.31. Their shapes are the same as those of the Wigner-
Seitz cells of the corresponding direct lattices since the Bravais types of the
direct and reciprocal surface lattices always coincide.
The first surface B Z s are, by definition, free of Bragg reflection lines.
Thus the energy function E(q) is continuous within these zones. Further-
more, any higher surface B Z of order p may be reduced to the first surface
B Z , and the energy function E(q) in the p-th surface B Z can be folded
back to the first surface B Z . There, it forms a continuous function E,(q)
which is called a surface energy band. Thus we may state that the energy
eigenvalues of a crystal with surface form energy bands over the first surface
BZ.
There are surface bands of different types, regarding their relations to
the energy bands of 3D bulk crystals without surface. Below we characterize
these differences in a qualitative way.

Types of eigenstates

Bulk states

Consider an infinite bulk crystal, and cleave it into two semi-infinite crystals
with a surface parallel to a particular lattice plane. The spectra of energy
3.6. Clesn semiconductor surfaces 359

t' tY

Figure 3.31: Surface B Z s of the 5 plane lattices: (a) oblique, (b) p-rectangular,
(c) c-rectangular, (d) square, (e) hexagonal. Symmetry lines and points are also
shown, and their notations are introduced.

eigenvalues of the two semi-infinite crystals will contain all energy levels
which were already eigenvalues of the infinite bulk crystal before cleaving.
This implies that each crystal with surface possesses an energy eigenvalue
spectrum which partially is made up from energy eigenvalues of the infinite
bulk crystals from which it is derived. The eigenfunctions of the crystal with
surface corresponding to these eigenvalues, if examined at positions outside
of the crystal, will decay exponentially with increasing distance from the
surface, while inside they will practically be the same as those of the infinite
crystal, i.e. they will exhibit undamped oscillations throughout the whole
semi-infinite crystal, as do the eigenfunctions of the infinite bulk crystal (see
Figure 3.32). Eigenstates of a crystal with surface exhibiting such properties
are called bulk states. The corresponding energy eigenvalues form bulk state
surface energy bands Ep(q). These bands can be obtained by projecting the
bulk bands E?"(k) of the infinite crystal onto the first surface B Z . Here,
'projecting' means that one assigns to a particular point q of the first surface
B Z all bulk band energies EF"(k) corresponding to k-vectors of the first
360 Chapter 3. Electronic structure of semiconductor crystals with perturbations

I
x C
f
.-0
P
W
-w
0
C C
3
w +
W

surface state

9- 0 x-

Figure 3.32: The three different types of electron energy eigenstates of a crystal
with surface below the vacuum level (schematically).

bulk B Z having the same projection q on the surface B Z plane, but with
various components kl perpendicular to it. Formally, this can be expressed
by the relations

The bulk state surface band index p in (3.213) replaces the bulk quantum
numbers v k l . As kl varies continuously, @ also does. This means that a
particular band of the infinite bulk crystal gives rise to a continuum of surface
bands. In this way, the infinitely large number of atoms in a primitive unit
cell of the crystal with surface manifests itself.
We illustrate the projection of bulk bands onto the surface B Z using the
(100) surface of a diamond type crystal as an example. The band structure
is taken from the empty lattice model introduced in section 2.4. As a first
step, the bulk B Z is to be projected onto the plane of the first surface B Z
(see Figure 3.32). In doing so, one notes that part of the projected bulk B Z
lies in the second surface B Z . This has to be folded back to the first surface
B Z , together with its energy values. In this way we obtain the projections
of the lowest three empty lattice bands shown in Figure 3.33. In addition
to q-vectors for which all energy values are allowed, there are also q-vectors
occurring for which certain energy values are forbidden. This peculiarity is
found in other, more realistic cases as well, it represents a general feature
of the projected bulk band structure. In the forbidden energy regions states
may occur which are localized at the surface.
3.6. Clean semiconductor surfaces 361

Figure 3.33: First (100) surface


B Z (shaded area) together with
I
the projected first bulk B Z for a
crystal with fcc Bravais lattice.

Surface states

Such states are to be expected for the same reason that localized states are
observed in the case of point or 0-dimensional perturbations. In contrast to
the latter, surfaces constitute 2-dimensional perturbations. The localization
occurs, therefore, only in 3 - 2 = 1 dimension, namely that perpendicular to
the surface. If the energy lies in the forbidden region of the projected bulk
band structure, the decay of the eigenfunction towards the bulk proceeds
exponentially (see Figure 3.32). The states are then called bound surface
states and the corresponding energy bands E,(q) are bound surface bands.
Besides these, one has surface resonances and antiresonances. The latter
occur at energies in the allowed region of the projected bulk band structure
and give rise to resonant or antiresonant surface bands Ep(q). They mani-
fest themselves in an increase (resonance) or decrease (antiresonance) of the
density of states. The eigenfunctions at these energies are also localized at
the surface, but decay less rapidly towards the bulk (according to a power
law) than the exponentially decaying bound surface states. Antiresonances
are necessary in order to satisfy Levinsons theorem, which holds for surfaces
as well as for point perturbations. If bound surface states exist, the antires-
onances must compensate the increase of the total DOS in the previously
forbidden energy region.

Implications of symmetry for surface band structure

The spatial symmetry of a crystal with surface has implications for the pos-
sible degrees of degeneracy of surface energy bands Ep(q), p = p, (T,p, at a
362 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.34: Projection of the


empty lattice band structure of
a fcc bulk crystal onto the first -u1l 9
(100) surface B Z . Three bulk
bands with reciprocal lattice vec-
5n
tors K = 0 (1 = 0), K = -2 7
( 2 n / a ) ( l l l ) ( I = 1) and K = %
E?
( 2 n / a ) ( m ) ( I = -1) are consid- a,
C
W
ered.
5

r K r

given wavevector q. It also results in symmetry relations between the values


of EF(q)for different q values, and it determines the spatial symmetries of
the eigenfunctions $i)w(~).The key for such conclusions are, in analogy to
the infinite bulk case, the irreducible representations of the space group of
the given crystal with surface. This is based on the fact that the eigenfunc-
tions for a particular energy eigenvalue form a basis set of an irreducible
representation of this group. Such a representation may be characterized
by the star {q} of the wavevector q and the irreducible representations of
the small point group of q with the factor system of equation (A. 157). The
dimensions of the irreducible representations determine the possible degrees
of degeneracy of an energy band E,(q) at the point q. Moreover, at all
points of the star {q}, E,(q) has the same value. Since the 10 possible point
groups of equivalent directions have at most 12 elements (this happens in
the case of C6v(6mm)),and since the small point groups of the symmetric
q-points are in general even smaller than the corresponding point groups of
equivalent directions, only irreducible representations with dimensions equal
to 1 or 2 will appear. 2D representations are likely to occur in cases where
the small point groups are, on the one hand, large enough, and on the other
hand, the corresponding space groups contain glide reflections. Then non-
trivial factor systems arise for points q on the boundary of the first BZ,
3.6. Clean semiconductor surfaces 363

pig1 r, -A;-?;--Z;- ~r~r+-~2---~-~;-X;--~;-M2-Z;- I x2

r, -A\-x:-z;--M,--z,- x,-A,- ~--A~;-&-z;-M,-zF- x1

I I I

Figure 3.35: Symmetry and degeneracy of the surface energy bands for the 5 space
groups of the p-rectangular Bravais lattice. [After Terzibaschian and Enderlein,
1986.)

and 1D representations might not be possible at all. This happens for 2 of


the 5 space groups of the primitive rectangular Bravais lattice, which below
will be studied as an example. In Figure 3.35, the possible types of band
structures are shown on certain symmetry lines of the p-rectangular surface
B Z . In the case of the space group p2mg, which applies for (110) surfaces of
diamond type crystals, only 2D representations exist at M and X . The two
1D representations on the 2-line connecting these points belong to the same
energy eigenvalue because of time reversal symmetry. For the space group
p2gg, one has 2D representations only at X and X'. The representations on
the connecting line X - 2 - M - 2' - X ' are lD, but the corresponding en-
ergy eigenvalues are degenerate because of time reversal symmetry. For the
three remaining space groups p2mm, p l m l , and p l g l , the representations
at all symmetry points are 1D.

Numerical methods for calculation of the electronic structure of


surfaces

A 3D crystal with surface may be characterized as a crystal with a 2D lattice


and primitive unit cells extending infinitely in the direction perpendicular
364 Chapter 3. Electronic structure of semiconductor crystals with perturbations

to the surface. This point of view gives rise to the so-called slab method for
calculating the electronic structure of crystals with surfaces.

Slab method

The electronic structure of a crystal with a 2D lattice may be calculated by


means of any of the methods known from band structure calculations of 3D
crystals, with the exception that the dimensionality of the lattice has to be
changed from 3 to 2. In the tight binding method, for example, one has to
use Bloch sums of atomic orbitals &(x) over all points r of the 2D surface
lattice, rather than over all points R of the 3D bulk lattice. With the surface
adapted notation, r+G+Zf3, for a regular crystal site, the orbital a localized
at such a site is given by &(x - r - 3 - 1f3) E &jl,.(x), The corresponding
Bloch sums &jlp(x)are defined as

(3.214)

The number of different Bloch sums is infinitely large even if a finite number
of orbitals is used per atom, because the integer 1 counting the primitive
crystal slabs, runs from 0 to -co. Consequently the Hamiltonian is given
by an infinite matrix in the atomic orbital representation &jl,.(x). Some
specific approximation is necessary to transform this to a finite matrix. One
possibility is to consider a slab of the crystal, i.e. to cut off the semi-infinite
crystal at a particular lattice plane parallel to the surface and discard the
remainder. One may say that, in the direction perpendicular to the surface,
the true semi-infinite crystal is replaced by a cluster (see Figure 3.36). The
latter has two plane surfaces, one of them being real (that of the semi-infinite
crystal), and the other one (obtained by cutting the semi-infinite crystal) not.
This differs from the cluster method in the case of point defects discussed in
section 3.5 where the whole cluster surface is artificial.
The slab method may be combined with any band structure calculational
method for infinite bulk crystals. Its combination with the tight binding
method will be illustrated below by means of an example.
We consider the ideal (111) surface of a diamond type crystal which
according to subsection 3.6.2, has a hexagonal Bravais lattice. Instead of
the one s- and three p-orbitals per atom we use the four hybrid-orbitals, i.e.
we replace a in the Bloch sum (3.214) by ht, t = 1 , 2 , 3 , 4 . Only nearest
neighbor interactions are taken into account. An illustration of this model
is given in Figure 3.37. To get the matrix elements of the Hamiltonian
H between Bloch sums, we first need these elements between the localized
hybrid orbitals +htjlT Ihtjlr). The diagonal elements are

(3.215)
3.6. Clean semiconductor surfaces 365

Figure 3.36: Illustration of the slab method for surface band structure calculations.

Figure 3 . 3 7 Defect molecule model of the ideal (111) surface of a crystal with
diamond structure.

and the non-diagonal elements between different hybrids at the same atom
j = j are
(htllrlH(htJ1lr)= VI , t # t . (3.216)

The two types of matrix elements in (3.215) and (3.216) are equal for atoms
at the surface and in the bulk. This is not true for elements between hybrids
located at different atoms. Let j = 1,1= 0, designate the surface atom layer
and consider the elements (ht10rJH]ht20rt)between the hybrid ht located
at the surface atom 10r and the hybrid lht20x-t) at its nearest neighbor 20rt
below the surface, pointing toward (htlOr). Since no nearest neighbor hybrid
exists for the hybrid )hllOr) pointing out of the surface, all nearest neighbor
elements involving IhllOr) vanish. The other three elements with t = 2 , 3 , 4
are equal to the parameter V2 introduced in equation (2.292),
366 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Using equations (3.215), (3.216) and (3.217), the matrix elements between
the corresponding Bloch sums (3.214) may be formed. They are

(htlOq(Hlht,lOq)= v1 , t # t . (3.219)

(htIOqlHlht20q) = etV2, t = 2,3,4 , (3.220)

where e t is given by equation (2.240), with dt being a vector which points


from the surface atom 10r toward its nearest neighbor atom 20rt in the
direction of the hybrid t. We have
a a - -
d2 = -(lTT), d3 = -(lll), d4 = a(TT1). (3.221)
2 2 2
The hybrids lht201-i) at the nearest neighbor atoms pointing td the surface
atom lor, are coupled to the other hybrids at these atoms, and these hybrids
interact with hybrids at more remote atoms. Thus an infinite matrix would
occur if we would not restrict consideration to a slab, as we in fact do.
Here, we will go one step further and neglect all couplings between the
hybrids at the nearest neighbor atoms. Then the 7 Bloch sums IhtlOq),
t = 1,2,3,4, and (ht20q),t = 2 , 3 , 4 , are completely decoupled from the rest
of the Bloch sums (see Figure 3.67). In this way we arrive at a simplified
model of the crystal with surface which effectively reduces it to the two first
atomic layers and treats the atoms of the second layer only in an approximate
way. This model represents an analog of the defect molecule model in the
case of a point perturbation. If the basis functions are arranged in the
order IhllOq), IhzlOq), h l o q ) , h l o q ) , lh220q), lh320q), lh420q), then the
Hamiltonian matrix is

(3.222)

The eigenvalues of this matrix are plotted in Figure 3.38 for different sym-
metry lines of the first surface B Z of the hexagonal lattice of Figure 3.31.
3.6. Clean semiconductor surfaces 367

r M K r
W avevec tor

Figure 3.38: Band structure of the ideal (111) surface of a diamond crystal within
the defect molecule model. The TB parameters are V1 = 2.13 e V , V2 = 6.98 e V .
The hybrid energy Eh is used as the energy origin.

The three lowest and three highest bands are bulk state surface bands. They
arise from the three hybrids of the surface atom, which bind this atom back
to its nearest neighbors in the second layer. More strictly, they correspond
to the 3 back bonding and antibonding states. The band in the gap is due
to the dangling hybrid of the surface atom. It forms a bound surface band.
If we were to consider more than 2 atomic layers, the region where the bulk
bands in Figure 3.38 occur would be covered by more bulk state surface
bands, while the bound surface band in the gap would hardly change.

Supercell method

The slab considered in the preceding subsection may be repeated periodi-


cally in the direction perpendicular to the surface, simultaneously inserting
several layers of vacancies between two neighboring slabs (see Figure 3.39).
This structure may be considered to be an infinite repetition of the origi-
nal crystal with surface, each repetition being approximated by a finite slab
of several atomic layers embedded between vacancy layers. This arrange-
ment represents a 3 D supercrystal composed of 1D supercells (bear in mind
that in the case of point defects in section 3.5, we similarly considered a
368 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.39: Supercrystal obtained by a periodic repetition of supercells. The


latter are composed of crystal slabs embedded between vacancy layers.

3D supercrystal composed of 3 D supercells). The band structure of the su-


percrystal is approximately the same as that of the original crystal with
surface, provided the vacancy slabs are thick enough to suppress coupling
between neighboring crystal slabs, and the latter slabs are thick enough to
simulate a semi-infinite crystal. The bands Ep(q) of the crystal with surface
are obtained from the bulk bands EF(k) of the supercrystal by plotting the
latter on the surface B Z . In this, k l may be chosen arbitrarily because
any dispersion of EF(k) = E F ( q , k l ) with respect to k l would indicate a
coupling between the slabs, which has been excluded. The band structure
of the supercrystal may be obtained by means of any 3 D band structure
calculation method without modification. This makes the supercell method
particularly appealing. Combined with the pseudopotential method, as well
as the local density functional or quasi-particle approximations, it represents
the most important calculational method for electronic and atomic structure
determinations of surfaces.

Defect model and Greens function method

A crystal with surface may also be viewed as a 3 D crystal with a 2D pertur-


bation. The significance of this characterization is the following: Consider
first an ideal infinite 3 D crystal, then remove some number of neighboring
atomic layers parallel to the surface under consideration or, equivalently,
3.6. Clean semiconductor surfaces 369

Figure 3.40: Illustration of the defect method for calculating surface energy band
structure.

create the same number of vacancy layers (see Figure 3.40). What remains
are two identical semi-infinite crystals which are only displaced with respect
to each other. They do not interact provided the number of vacancy layers
is large enough, which we will assume. The electronic structures of the two
semi-infinite crystals coincide, and are identical with that of the considered
crystal with surface. As in the case of a single vacancy, the states bound at
the defect may be obtained by means of the Greens function G o ( E )of the
unperturbed 3D crystal, more strictly, by the vanishing of the determinant
of [Go(E)V - 11 (see equation (3.134)). Here, the perturbation potential
V(x) represents the difference of the potential energy of an electron in the
perturbed and unperturbed crystals. It is the negative of the sum of the
potentials of the removed atoms or the sum of all vacancy potentials. It has
the 2D lattice symmetry of the surface so that

~ ( x=) V(X + r). (3.223)

If only nearest neighbor interactions are taken into account, a single layer
of vacancies is enough to decouple the two semi-infinite crystals. For the
analysis of the bound state condition (3.74), one has to use a particular basis
set, just as in the case of a point perturbation. Wannier functions are again
a possible choice, but here they involve localization only in one direction,
namely that perpendicular to the surface. We denote these functions by
IvqRl) where RI is the component of a 3D lattice vector R perpendicular
to the surface, such that
370 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.224)

The matrix representation of the Greens function Go(E) with respect to


this basis set is

For the perturbation potential V1(x) we take that of a single layer of vacan-
cies located at RI = 0. Then the matrix representation of V1(x) is

where VL,,(q) = (vq01VJvqO)has been used. The relevant matrix elements


(vqOIGoV1- 1lvlqO) of GoV1- 1 take the form

with
(3.228)

According to equation (3.134), the determinant of the matrix (3.227) must


vanish for energy eigenvalues E in the gap of the ideal crystal, i.e.

Det [G:(E, q)VL,,(q) - 6,,,] = 0. (3.229)

If one compares this equation with the corresponding relation (3.94) for deep
centers, it will be noted that the q-dependence which occurs here was absent
there. This dependence in the case of surfaces causes the eigenvalues in the
gap to form deep bands rather than deep levels (which was the case for
point perturbat ions).

Transfer matrix method

Finally, still another method for surface band structure calculations should
be mentioned. It is based on the transfer matrix concept of quantum mechan-
ics. The transfer matrix M ( E ) is formed from solutions of the Schrodinger
equation upon one unit cell of the bulk crystal for particular boundary condi-
tions in the direction perpendicular to the surface. The energy E is arbitrary
first of all. Transferring the wave function from the surface unit cell to the
3.6. Clean semiconductor surfaces 371

n - th unit cell below the surface can be done by applying the n - t h power
W ( E ) of M ( E ) . Bound surface states are obtained for such energy values
E for which M n ( E ) decays exponentially with n. The practical use of this
method seems to be limited, however.

3.6.4 A t o m i c and e l e c t r o n i c structure of particular surfaces

In this subsection we will deal with the atomic and electronic structures of
some important semiconductor surfaces, including the various reconstruction
states of the Si (111) surface, the Si (100) surface, the (110) surface of GaAs
and other 111-V compound semiconductors as well as the (111) and (100)
surfaces of GaAs. It is advisable to treat the atomic structure together with
the electronic structure because of the close relation between the two kinds
of structures, as was pointed out earlier. We begin with a short introduction
to the experimental methods of surface structure analysis, again referring to
both the atomic and electronic aspects.

Experimental methods for surface s t r u c t u r e analysis

Experimental methods for determining the atomic structure of bulk crystals


are all based on the interaction of waves with the atomic cores and valence
electrons of the crystal. If the wavelength is of the order of the distance
between the atoms, i.e. of the order of magnitude of 1 A, the crystal consti-
tutes a 3D diffraction lattice and diffraction maxima will occur in prescribed
directions in space. The crystal structure may be determined from the po-
sitions and intensities of these maxima. The various experimental methods
differ primarily in the nature of the waves employed. X-rays are by far the
most important for determining the structure of bulk crystals. A wavelength
in the region of 1 corresponds to a photon energy in the range of 10 k e V .
Electron and neutron waves are also diffracted by bulk crystals. Electron
energies in the region of 100 eV and neutron energies of 0.1 e V are required
for wavelengths in the d region. Since they are neutral particles, X-ray pho-
tons and neutrons interact only relatively weakly with the crystal. They can
pass through crystals of macroscopic thickness and be backscattered from
them from macroscopic depths within them. X-rays and neutrons thus yield
information on all atomic layers of a crystal, including those at the surface.
Since the number of surface layers is extremely small in comparison to the
total number of layers, the diffraction patterns are dominated by the bulk
of the crystals.
The interaction of electrons with the atomic cores and the valence elec-
trons of a crystal is significantly stronger than that of photons and neutrons.
Electrons having energy less than 100 k e V can not pass through a crystal of
372 Chapter 3. Electronic structure of semiconductor crystals with perturbations

macroscopic thickness. Experimentally, one therefore has only the backscat-


tering available and then only elastically backscattered electrons can be em-
ployed in forming diffraction patterns. These originate at a depth which,
on average, is equal to the inelastic mean free path of the electrons. This
varies relatively independently of the particular crystal under consideration
from 4 A to 10 for energies of 20 eV to 300 eV. Electron diffraction in
this energy region is thus hardly suitable for examining bulk crystals but it
can be readily employed in studying crystal surfaces. The diffraction of low
energy electrons in the region of 100 eV is in fact the most intensively used
method for surface structure determination. It is referred to as LEED (Low
Energy Electron Diffraction).
The principle of LEED may be explained as follows. We consider an
incident electron wave with the wavevector ki. The interaction with the
crystal generates scattered waves with wavevectors k,. The scattering po-
tential has the translation symmetry of the surface lattice, thus its Fourier
components differ from zero only for vectors g of the reciprocal surface lat-
tice. This means that only those scattered wavevectors k, can occur whose
components k,ll parallel to the surface differ from the parallel component
kill of ki by a reciprocal surface lattice vector g, hence

(3.230)

There is no relation between the components of k, and ki perpendicular


to the surface because there is no translation symmetry of the scattering
potential in this direction. In writing down equation (3.230) we have implic-
itly assumed that only one scattering event takes place. This relation also
applies, however, to multiple scattering processes. This is important be-
cause electrons scattered back from the surface have, as a rule, experienced
many scattering events, in contrast to X-ray photons which typically have
been scattered only once. This difference is due to the above mentioned fact
that electrons interact with the crystal much more strongly than do X-ray
photons. The electrons measured in LEED are elastically scattered. One
therefore has

Ik,/= lkzl . (3,231)

A solution of the two equations (3.230), (3.231) always exists for given vec-
tors ki and g (this is in remarkable contrast to coherent scattering of elec-
trons from 3D bulk crystal, which can only occur if k, lies on a Bragg re-
flection plane). The solution of equations (3.230) and (3.231) can be readily
carried out using the construction shown in Figure 3.41. The points at which
the vertical lines passing through the reciprocal lattice points g intersect the
3.6. Clean semiconductor surfaces 373

Figure 3.41: Construction of


LEED maxima.

sphere Jk,]= lkzl, determine the directions in which diffraction maxima oc-
cur. There is exactly one maximum for each reciprocal lattice point g. The
reciprocal surface lattice can thus be read immediately from the distribution
of the diffraction maxima on the registration screen. The direct surface lat-
tice is the reciprocal of the reciprocal surface lattice. Some typical LEED
images are shown in Figure 3.42. The bright points correspond to the re-
ciprocal lattice of the ideal surface, and the less bright points to the finer
reciprocal lattice of the reconstructed surface. In this way it is relatively
easy to determine the surface lattice by means of LEED. To obtain the ac-
tual positions of atoms is more diacult. One needs additional experimental
and theoretical information about the intensity of the diffraction maxima as
a function of the energy of the incident electrons (dynamical LEED).
Besides LEED, there are other methods for surface structure analysis
which, although they are not a substitute for LEED, can supplement it.
These methods include diffraction of energetic electrons in the region of some
10 k e V , known as Re5ection High Energy Electron Diffraction ( M E E D ) ,
diffraction of X-rays incident almost parallel to the surface, and diffrac-
tion of slow Helium atoms (of M I00 meV). Scattering of energetic ions
( M 1 M e V ) is used in techniques like Rutherford backscattering (RBS) and
ion channeling. Imaging procedures of significance are transmission elec-
tron microscopy (TEM), scanning tunneling microscopy (STM) and atomic
force microscopy (AFM). The latter two methods have become particularly
important.
Like for atomic structure determinations, a variety of methods exist to
study the electronic structure of surfaces, in particular the bound surface
states in the energy gap of the bulk crystal. The most powerful and uni-
versal method is photoemission spectroscopy (PES). This method relies on
the external photoeffect in which an electron is emitted from the crystal by
374 Chapter 3. Electronic structure of semiconductor crystals with perturbations

x6

Figure 3.42: LEED pictures of six differently prepared GaAs (100) surfaces.(After
Drathen, Ranke and Jacobi, 1978.)

absorbing a photon of sufficiently high energy. The emitted photoelectrons


are spectrally decomposed with respect to their kinetic energies. The thus
obtained energy spectrum of photoelectrons maps the density of states of
occupied electron levels of the crystal. To enhance surface states and dis-
criminate bulk states, photoelectrons with kinetic energies around 50 eV
are used whose inelastic mean free path is only about 5 A and which can
therefore only come from this depth below the surface. These electron ener-
gies correspond to photon energies which are not substantially larger, i.e. in
the far ultraviolet region. The term UPS (Ultraviolet Photoemission Spec-
troscopy) is used in this context. The only practically suitable radiation
source in this energy region is the electron synchrotron. By measuring an-
gular resolved photoemission spectra (ARUPS), the wavevector dispersion
of the bound surface energy bands can be determined. To study moccu-
3.6. Clean semiconductor surfaces 375

11101 hi01

Figure 3.43: Geometry of the ideal Si (111) surface (left) and of the Si (111) 2 x 1
surface according to the buckling model (right).

pied surface states one may use inverse PES in which electrons captured
by such states emit photons. Beside photoemission, a variety of other tech-
niques exists which can provide data on surface states. In principle, any
experimental technique which probes the electronic structure of bulk crys-
tals can be employed for surface electronic structure investigations, provided
it can be made surface-sensitive. This applies to optical reflectivity, elec-
trical transport, photoconductivity, and capacity measurements, as well as
electron energy loss spectroscopy (EELS). Experimental techniques like field
effect measurements fulfill this requirement from the very beginning. Con-
trolling the energy of tunneling electrons in scanning tunneling microscopy,
surface states can be resolved spatially and energetically (scanning tunneling
spectroscopy).
Experimental techniques which primarily measure the electronic struc-
ture, can also provide data on the atomic structure. The solid state shifts
of core levels (see section 2.1) are an example. These shifts differ for atoms
in the bulk and at the surface because of the altered atomic structure at
the surface. The difference (typically some tenths of an e v ) can be mea-
sured by means of PES and UPS. On the other hand, they can be calculated
on the basis of a particular surface structure model. By comparing theory
and experiment one can evaluate the feasibility of various models of surface
structure.
The calculation of the total energy is a purely theoretical test of the
validity of a particular surface structure model, and it may be used to deter-
mine the parameters which can be varied in such a model. If the model has
optimized parameters and results in a lower total energy than other models
it may be given preference over them.
376 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.44: Surface band


structure of an ideal Si (111) sur- I ' O r
face.

t i
*!:!

!ii
::I .....
-5 - *
.. .*....
..
..
,-I I I...
ll
iii

r I K r
2D wave vector

Si surfaces

(111) surface

The geometry of the ideal (111) surface of diamond type crystals is illus-
trated in Figure 3.43 (left). The surface lattice is hexagonal, and the two
primitive lattice vectors are fi, f2 of Table 3.8. There is one surface atom per
primitive unit cell, and one dangling bond per surface atom. The (111) sur-
face is the cleavage plane of diamond type crystals. By cleaving a Si crystal
in UHV at room temperature, one obtains a 2 x 1 reconstructed (111) sur-
face. After annealing it at 500 C , a 7 x 7 reconstruction state evolves, which
remains stable at room temperature. The occurrence of a 2 x 1 reconstruc-
tion immediately after cleavage is to be expected if one examines the band
structure of the ideal (111) surface (see Figure 3.44) and considers, in partic-
ular, the electron occupancy of the bound surface band in the fundamental
gap. This band arises from dangling sp3-hybrids of surface atoms, and can
host 2 electrons per surface unit cell. Since 3 of the 4 valence electrons of
a surface atom are in bonds with second-layer atoms, only 1 electron per
surface unit cell is left for the bound surface band. Thus this band remains
only half-filled. The ideal Si (111) surface is metallic.
3.6. Clean semiconductor surfaces 377

This state is unlikely to be stable, however, i.e. surface reconstruction is


likely to take place. Below we discuss a particularly simple reconstruction
model, the so-called buckling model (see Figure 3.43, right-hand side) which
in the early days of clean semiconductor surface physics was believed to be
correct. Later, it was realized that buckling is energetically advantageous
only for 111-V compound semiconductor surfaces, while it is not advanta-
geous for group-IV semiconductor surfaces including the (111) surface of
Si. To introduce the buckling model we consider doubling of the primitive
unit cell in the direction of primitive lattice vector fi, which according to
Table 3.8 points in the direction [ l i O ] . The hexagonal lattice with doubled
primitive unit cell forms a rectangular lattice with primitive lattice vectors
2f1 -tf2 and f2 The short side of the rectangular primitive unit cell, shown in
Figure 3.43 right, is parallel to [Olq,and its long side parallel to [ 2 m . The
corresponding first surface BZ is also a rectangle (see Figure 3.31) with its
long side, i.e. its r - X-direction, parallel to [Olq,and its short side, i.e its
X - M-direction, parallel to [2ii].The rectangular BZ is half as big as the
original hexagonal BZ, and each band of the latter gives rise two band in the
former, a direct and a back-folded one. There is no gap between these two
bands because they arise from the same band of the larger BZ. The surface
is still metallic.
A gap arises if the so far formal 2 x 1 reconstruction is made real. This
can be done by a buckling of the surface, i.e. by alternately raising and
lowering atoms in rows parallel to f1 above the surface and below it (see Fig-
ure 3.43, right). In this, the three back-bonding hybrids of a raised atoms
becomes more p-like, and the dangling hybrid at this atom more s-like si-
multaneously lowering its energy, while the three back-bonding hybrids at
a lowered atom become more sp2-like and the dangling hybrid at this atom
more plike simultaneously raising its energy. The two bound surface bands
derived from these s- and p-like dangling hybrids are just the bands below
and above the gap discussed before. The lower s-like band can host all elec-
trons of the dangling hybrids, while no electrons remain for the population
of the upper p-like band. If the total energy of this state were in fact lower
than that of the ideal surface, buckling would take place spontaneously, i.e
the translation symmetry of the surface would spontaneously be lowered,
A similar spontaneous symmetry breaking, the Jahn-Teller effect, was dis-
cussed in the context of point perturbations in section 3.5. There, the point
symmetry was broken, while no translation symmetry was involved. If the
translation symmetry is broken, as in the case of surface reconstruction, one
speaks of a Peierls instability or a Peierls t r a n s i t i o n
However, as has been indicated at the outset, buckling turns out to be
energetically not favorable in the case of Si (111) surfaces. Populating the
lower s-like band with two electrons per primitive surface unit cell means
transferring charge from the atoms lowered below the surface to the atoms
378 Chapter 3. Electronic structure of semiconductor crystals with perturbations

r?

'P
Ur-

[1?01 [I101

Side view

a) b)

Figure 3.45: a-bonded chain model of the Si (111) 2 x 1 surface (After Pandey,
1982). Part (a) shows the unreconstructed surface in top and side views. The top
view in the second row has been rotated with respect to the top view in the first
row in order t o allow for the side view below. Part (b) shows the same views of the
surface as in part (a), but after reconstruction has taken place.
3.6. Clean semiconductor surfaces 379

raised above. This implies the creation of an electric dipole which is too
costly in energy to actually take place. Using the terminology of section
2.2 we may say that correlation effects of electron electron interaction, more
strictly speaking, the configuration dependence of oneelectron states, pre-
vents the buckled Si (111) surface to be lower in energy than the ideal one.
The reconstruction model which actually applies to the Si (111) surface
is the so-called s-bonded chain model, illustrated in Figure 3.45. In this
+
model, second layer atoms in rows parallel to fi f2, i.e. along the [lo3
direction in Figure 3.45a (including atom number 2) are raised into the first
layer as shown in Figure 3.45b, breaking their bonds with atoms in the third
layer (for example, the 2-5 bond). The dangling bond of the new surface
atom (say atom 2) is used to establish bonds with atoms of the first layer
(the 2-1 bond in this case). These can only be s-bonds (indicated by double
lines in Figure 3.45b) because the dangling bonds are perpendicular to the
surface. In this way s-bonded chains occur along the [ l O q direction (Pandey,
1982). The dangling bonds left at the third layer atoms (for example, atom
5) are saturated by hybrids of the first layer atoms which have been lowered
down to the second layer (for example, atom 3). The surface is in fact 2 x 1
reconstructed. This may be seen by taking the primitive lattice vectors of
the ideal surface to be f 1 + f 2 and -f2. Doubling -f2 yields the rectangular
lattice indicated in Figure 3.45b by dashed lines. Its primitive lattice vectors
are f1 + f 2 and - 2 f 2 + (fi + f 2 ) = f 1 - f 2 so that the short side of the rectangle
is parallel to the chain direction [lOT], and the long side perpendicular to it
(parallel to [121]).A peculiarity of the n-bonded chain model is that it has
a different bonding topology in comparison with the ideal (111) surface and
also with respect to the buckling model. While the latter exhibit rings with
6 mutually bonded atoms (see Figure 3.45a), the former shows alternating
rings with 5 and 7 bonded atoms (Figure 3.45b). This is due to the fact that
bonds existing at the ideal and buckled surfaces are broken and new bonds
are established in the s-bonded chain model.
The total energy of this model is clearly below that of the ideal surface
(about 0.5 eV per surface atom). Thus it represents a good candidate for
the reconstruction of the (111) Si surface. Further evidence is provided by
ARUPS and optical measurements. Figure 3.46 shows the wavevector dis-
persion of the two bound surface bands as obtained from ARUPS measure-
ments together with the calculated dispersion of these bands. The agreement
is quite satisfying. The strong dispersion of the bound surface band on the
r-X-line and the weak dispersion on the X-M-line of the rectangular sur-
face B Z is easily understandable: the long r-X-side of the rectangular unit
cell in q-space corresponds to the short side of the rectangular unit cell in
coordinate space, which is also the direction of the s-bonded chains. One
expects strong dispersion along the chains and weak for the perpendicular
direction, exactly what is seen in Figure 3.46.
380 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.46: Dispersion of the


bonding (B) and antibonding
(A) bound surface state bands
along the r - X - M line for the
-t2
Si (111) 2 x 1 surface. Curves are
-2>
W
calculated within the 7r-bonding
chain model, points are obtained 3 1
0
d

by means from ARUPS measure- Q,


a
ments. (After Martensson, Cri- x
centi, and Hansson, 1985.) Yl
g o
W

-1

r X nf
20 wave vector

Figure 3.47: Differential reflectiv-


ity spectrum of the Si (111) 2 x 1
0
surface (After Chiarotti, Nannarone,
Pastore and Chiaradia, 1971.)
93 0.A $5 0,s
Energy (ev) - [ 7
3.6. Clean semiconductor surfaces 38 1

Figure 3.48: Polarization dependence of the differential reflectivity spectrum of


the Si (111) 2 x 1 surface of Figure 3.45 (taken at its maximum). The solid curve
is calculated using the x-bonded chain model, the dashed curve using the buckling
model, and the points are experimental data (After Del Sole and Selloni, 1984.)

Support for the r-bonded chain model comes also from optical measure-
ments. The differential reflectivity spectrum of the 2 x 1 reconstructed (111)
surface of Si is shown in Figure 3.47. There is no doubt that the observed
peak at 0.5 eV is due to optical transitions between occupied and unoccu-
pied bound surface bands. Such bands exist both in the buckling model
as well as in the 7r-bonding chain model (in the latter one has 7r-bonding
and x-antibonding bound surface bands). However, the two models differ in
regard to their predictions on polarization dependence of optical reflectiv-
ity. According to the 7r-bonded chain model, transitions with light polarized
parallel to the chains, i.e. parallel to [lOq, should be allowed and transi-
tions for light polarized parallel to the perpendicular direction [121]should
be forbidden, while this should be reversed for the buckling model. The ex-
perimentally observed polarization dependence shown in Figure 3.48 is that
predicted by .I-bonding chain model. It rules out the buckling model.

Besides the 2 x 1 reconstruction, there are other reconstruction states


of the Si (111) surface. The 7 x 7 reconstructed surface is the most stable
one. The complicated structure of this surface has finally been resolved by
combining the results of various experimental methods including STM (see
Figure 3.49). The model which accounts for all experimental data utilizes
three structural disturbances of the ideal surface, these being dimers (D),
adatoms (A) and stacking faults (S). It is referred to as the DAS model
382 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.49: Scanning tunneling microscopy (STM) image of the Si (111) 7 x 7


surface. (After Quate, 1986.)

(Takayanagi, 1984). The DAS model of the 7 x 7 reconstructed (111) Si


surface is shown in Figure 3.50.

(100)surface

The (100) surface is the surface of choice for electronic device applications
of Si. The geometry of the ideal (100) surface is shown in Figure 3.51. The
surface lattice is a square one, with primitive lattice vectors f 1 , f2 given in
Table 3.8. The primitive unit cell has one surface atom, and each atom
has two dangling bonds which point out of the surface like rabbit ears (see
Figure 3.51a).
Each dangling bond is only half-filled as in the case of the (111)surface
considered above. Thus, the ideal surface is metallic, and this state is un-
likely to be stable. A state of lower total energy can be established by a
2 x 1 reconstruction as follows: The atoms of two neighboring rows parallel
to [ O l i ] (or [Oll]) move slightly towards each other in order to allow bonding
between two of their four dangling hybrids. This dimerization of the sur-
face gives rise to bonding and antibonding bound surface states, the lower
bonding state being completely filled and the upper antibonding state being
completely empty. The primitive lattice vector in the direction of f 1 doubles,
thus a 2 x 1 reconstruction takes place, and the surface lattice becomes rect-
angular. The remaining two dangling bonds of a dimer are still half-filled,
however, so that the surface is not yet stable. It is stabilized by buckling,
3.6. Clean semiconductor surfaces 383

a1

Figure 3.50: Dimer-Adatom-Stacking-Fault (DAS) model of the Si (111) 7 x 7


surface. The side view (a) is shown to identify the atoms: large shaded circles
are adatoms, open circles are surface atoms of the first (large circles) and second
(smaller circles) monolayer, solid circles are bulk atoms which do not undergo re-
construction. The top view (b) shows a 7 x 7 surface unit cell and its surroundings.
The small circles within shaded circles represent second layer atoms vertically below
the adatoms. (After Takayanagi, 1984.)

which turns out to be energetically favorable in this case. One of the two
atoms of a dimer moves above the surface and one below. The dangling
hybrid at the lowered atom, being p-like, takes a higher energy and is corre-
spondingly empty, while the dangling hybrid at the raised atom, being s-like,
takes a lower energy and is filled. Rigorous structure calculations essentially
confirm this simple tight binding picture. They only add displacements of
the dimer atoms parallel to the surface in addition to the perpendicular ones.
Due t o the two kinds of displacements, the dimers become asymmetric. The
described 2 x 1 reconstruction of the Si (100) surface is therefore called the
384 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Top view (100)

I-- 0

0'- -0
n
It-
-?
0
U

10113

Side view

a) b)
Figure 3.51: Geometry of the ideal Si (100) surface (a), and of the asymmetric
dimer model of this surface.

asymmetric dimer model.


Although the 2 x 1 reconstruction is the most common superstructure of
the Si (100) surface, other reconstructions are also observed, for example,
2 x 2, c - 2 x 2, c - 4 x 2. Most of these structures may be traced back to
asymmetric dimers as building blocks.

GaAs and other 111-V compounds

(110) surfaces

The (110) surface represents the cleavage plane of zincblende type crystals.
The best understood (110) surface of all zincblende type 111-V semiconduc-
tors is the (110) surface of GaAs. Its surface geometry is shown in Figure
3.6. Clean semiconductor surfaces 385

Top view (110) Ga o As 0

I?
I
I
I
I
I
I
5
0
I
1
I
I
I 1 Y
I I
I I I I
I I

[?I0 1

Side view

-
0
c-
7
U

[?I01
a) b)

Figure 3.52: Geometry of the ideal GaAs (110) surface (a), and of the same surface
after relaxation (b).

3.52. The surface lattice is p-rectangular. The primitive lattice vectors are
given in Table 3.8. There are two surface atoms in a primitive unit cell, one
Ga and one As atom. Two bonds of each surface atom lie within the surface,
one is directed back and one is dangling. The two dangling hybrids per unit
cell have different energies since they belong to either a Ga- or an As atom.
Thus one expects two bound surface bands, one Ga-like and one As-like.
These are in fact seen in the band structure of the ideal (110) GaAs surface
depicted in Figure 3.53. The lower As-like bound surface band is completely
occupied, and the higher Ga-like band is completely empty. Thus the ideal
surface is semiconducting. Nevertheless, it does not yet represent the stable
state, as the gap between the two bound surface states is too small. It can
be enlarged by moving the As atom above the surface, rendering its dangling
hybrid more s-like and lowering its energy, and moving the Ga atom below
386 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Ga As
(111) (1101 (1001

-10

r M K r r x ' M X r r i ~ l r
2 D wave vector

Figure 3.53: Band structure of the ideal GaAs (110) surface. (After Talwar and
Ting, 1992.)

the surface, rendering its dangling hybrid plike and raising its energy (see
Figure 3.52). These displacements do not change the translation symmetry
of the surface because the two atoms belong to the same unit cell. One
therefore has a relaxation instead of a reconstruction of the (110) surface.
This is consistent with LEED measurements, which do not show spots other
than those due to the ideal prectangular lattice. The experimental value
for the rotation angle w of a Ga-As bond with respect to the [ l i O ] direction
is close to 30, in good agreement with structure calculations. Experiment
and theory also agree in regard to an essential feature of the band structure
of the relaxed (110) surface. As indicated in Figure 3.54, relaxation moves
the bound surface bands completely out of the fundamental gap, the As-like
band merges into the valence band, and the Ga-like band merges into the
conduction band. This implies that perturbations of the surface such as,
for example, coverage by an insulator, can easily create surface states in the
gap. Such states are in fact present at GaAs/insulator interfaces. They pin
the Fermi level and preclude the possibility of making GaAs-based field ef-
fect transistors in the same way as the Si-based MISFET (see Chapter 7 for
further discussion).
Relaxations similar to that of the (110) surface of GaAs are also observed
at the (110) surfaces of other 111-Vcompounds. For all materials except Gap,
relaxation moves the bound surface states out of the gap (see Figure 3.54).
3.6. Clean semiconductor surfaces 387

(110)cleavage face
GaP GaAs GaSb In P In As In Sb
- A
2
-
h

-3-
-0
5 -4-
3
u
9 -5- U n
g
3 -6-
x
e
E -7-
w

Figure 3.54: Energy level diagrams of bound surface states and bulk states of var-
ious 111-Vcompound semiconductors below the vacuum level. Solid lines represent
experimental results, and dashed regions represent dangling bond surface bands.
F'ramed undashed regions indicate bulk bands. (After Bertoni, Bisi, Calandra, and
Manghi, 1978.)

Other surfaces

Besides the (110) cleavage surface, investigations have mainly been focused
on the low index (111)and (100) surfaces of GaAs. (100) is the preferred sur-
face orientation of GaAs wafers used in device fabrication (partially because
the (011) plane perpendicular to this surface represents the cleavage plane
of GaAs). The geometrical structures of the ideal surfaces are the same as
those of the corresponding Si surfaces shown, respectively, in Figures 3.43
and 3.51. The (111)and (100) surfaces of GaAs differ from the (110) surface
of GaAs mainly in regard to the fact that two different surface terminations
are possible in their case, one by Ga atoms and another by As atoms. One
says that these surfaces are polar, in contrast to the (110) surface, which is
said to be non-polar. In forming a polar surface, an electric dipole is created
between a Ga-layer and an As-layer which is costly in energy. This explains
why for GaAs and other 111-V compound semiconductors, the cleavage plane
is neither ( l l l ) , like in the case of group-IV materials, nor (loo), but the
non-polar (110) plane. In the latter case the 1:l ratio of Ga- and As atoms
is strictly k e d by chemical stoichiometry. For the polar surfaces, there are,
however, no stoichiometrical reasons which cause a surface termination by
388 Chapter 3. Electronic structure of semiconductor crystals with perturbations

only Ga- or only As atoms. Both kinds of terminations can occur simulta-
neously at a given surface, and in order to define the surface uniquely one
has to specify the percentage of Ga and As it contains. A rough distinction
is that between Ga-rich and As-rich surfaces.
The structure of a particular polar GaAs surface depends decisively on its
termination. For the Ga-rich (111) surface one finds a 2 x 2 reconstruction,
and for the As-rich case there are f i x 8- 30 and a x - 23.4''
reconstructions in addition. For the Ga-terminated 2 x 2 (111) surface, a
model has been proposed with a quarter of the surface atoms missing. The
remaining first-layer Ga atoms and the second layer As atoms undergo a
buckling, which raises the As atoms close to the surface.
The structure of the GaAs (100) surface exhibits an even greater vari-
ety, depending on surface termination, surface treatment and temperature.
Some of the possible reconstruction states can be seen in the LEED pic-
tures of Figure 3.42. The c - 2 x 8 structure is found for an As-stabilized
surface, which is important for MBE-growth because this commonly begins
and ends with As-rich conditions. For the Ga-stabilized surface, the c - 8 x 2
reconstruction is found to be stable.

3.7 Semiconductor microstructures

Semiconductors with a clean planar surface, considered in the section above,


may be thought of as units of two infinite half spaces, one filled with the
semiconductor material, and the other being empty. If the vacuum is re-
placed by a semiconductor material different from the first, then one obtains
a semiconductor heterojunction or single semiconductor heterostructure. Fig-
ure 3.55 shows an example. Below the plane at z = 0 one has semiconductor
material 1, say GaAs, and above it is the material 2, say AlAs. The plane
at z = 0 is called the interface. The microstructures to be considered in this
subsection are composed of semiconductor heterojunctions. Thus, before
addressing microstructures we must deal with het eroj unct ions.

3.7.1 Heterojunctions
Below we describe the electronic structure of semiconductor heterojunctions.
The two semiconductor materials are taken to be undoped. Free carrier ef-
fects on the electronic structure can be neglected in these circumstances.
In the case of heterostructures formed from doped semiconductors, such ef-
fects might be important. They are treated in Chapter 6 in a systematic
way. Below, we discuss the various stationary one-electron states of undoped
heterostructures, omitting free carrier effects.
3.7. Semiconductor microstructures 389

Stationary one-electron states

As in the case of a crystal with a clean surface, heterojunctions of the kind


described above possess a 2-dimensional rather than a 3-dimensional lattice
translation symmetry. Generally, their one-electron states cp(x) are Bloch
states (Pk,,(xI(,z ) in regard to their dependence on the position vector com-
ponent XIIparallel to the interface, with 2-dimensional quasi-wavevectors kll.
The corresponding energy eigenvalues form bands in the 2-dimensional first
B Z of the heterostructure. Just as in the surface case, almost all energy
eigenvalues of the two infinite bulk materials, i.e. the Bloch bands E,l(k)
and E,a(k) of these materials, are also energy eigenvalues of the heterojunc-
tion. What may change are the eigenfunctions of these energy bands. Let us
fix a particular quasi-wavevector ko. If an energy level E,n(ko) of material 2
does not coincide with any of the allowed energy levels E,l(ko) at ko of ma-
terial 1, then the eigenfunction belonging to this level and quasi-wavevector
will be localized in material 2. There, its wavefunction is spread out uni-
formly over the whole semi-infinite crystal from z = 0 to z = +co;it forms
a bulk state of material 2. This is illustrated in Figure 3.56a by representing
energy levels of this kind by lines extending from z = 0 to z = +co. Vice
versa, if, at ko, an energy level E,l(ko) of material 1 does not coincide with
any of the allowed energy levels E,z(ko) of material 2, it forms a bulk state
of material 1, and may be represented by a line extending from z = -co to
0. If there are energy levels at ko which are identical in both materials, i.e.
with E,l(ko) = EA(ko), the corresponding wavefunctions will extend over
both materials (Figure 3.56b). In the case of identical energy levels E,l(ko)
and E,a(kb) at dzflerent wavevectors ko and kb, it is not a priori clear what
will happen. If a matching of the corresponding two eigenfunctions and their
derivatives at the interface turns out to be possible, then states extending
over the whole heterojunction from -cc to f m will exist (Figure 3.56~).If
390 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Material 1 I Material 2

P
a,
L.

Figure 3.56: Various bulk states of a semiconductor heterojunction.

no matching is possible, then the two eigenfunctions will remain localized in


their respective material regions. One has a situation similar to that in the
case of electromagnetic waves propagating between two semi-infinite dielec-
tric media: for certain wavevectors they may propagate from one medium
into the other, and for others they are internally reflected, as indicated in
Figure 3 . 5 6 ~ .
Besides bulk states, there may be stationary states with energy eigenval-
ues which are allowed in none of the two materials. The wavefunctions of
such states are localized at the interface between the two materials, just like
the bound surface states in the case of clean surfaces. They are called bound
interjuce states. In analogy to the clean surface case, interface resonances
3.7. Semiconductor microstructures 391

may also occur, their energy levels lie in one of the bands and their wave
functions are weakly localized at the interface. If the two materials forming
a heterojunction are composed of chemically similar atoms as in the case
of GaAs and AlAs and in many other cases of practical importance, then
no bound interface states will be possible, because the perturbation at the
interface is too weak.

Valence band discontinuity

In many semiconductors, the maximum energy of the valence band occurs


at the center r of the first BZ. Below we will restrict ourselves to materials
of this kind. Of course, the maximum energy of the valence band itself
depends on the material under consideration. This statement may seem to
contradict the results of Chapter 2, where this energy was set to zero for
all materials. However, this was done in the context of dealing with the
band structure of only o n e infinite bulk semiconductor at a time. For such
a single semiconductor, the energy origin could be chosen arbitrarily, and
we took it to be at the valence band maximum. In the heterojunction of
Figure 3.55, two semiconductor materials are involved, but the energy origin
of the heterojunction can be fmed only once. If the valence band maximum
is taken as zero for one material, it will, in general, differ from zero for the
other material. Here, instead of setting it to zero for any material, we select
the vacuum level as the common energy origin. This level may be defined as
the minimum energy which an electron in an infinite semiconductor sample
must have in order to escape. We will denote the valence band edge of a
particular material i , i = 1,2, referred to the vacuum level as origin, by E$.
Then the minimum energy which must be expended to remove a valence
electron from material i , is given by --E$ If the electron should escape
by absorbing a photon, then -Eti is the minimum photon energy required
(photo-t hreshold energy).
For the actual positions of the valence band edges Evl and Ev2 at a
heterojunction, the infinite bulk values Etl and Et2 have only an indirect
meaning. In fact, for each of the two semiconductors of a heterojunction the
adjacent material is foreign and represents an external perturbation. Even if
no free charge carriers are available, as we assume, each of the semiconduc-
tors reacts to this perturbation by redistributing its electrons, in this case
its bound valence band electrons. How this occurs is quite clear physically,
and it can also be analyzed in a more rigorous treatment: valence electron
charge will flow from the material with the higher valence band edge into the
material with the lower valence band edge, thus lowering the total energy
of the heterojunction. In this way, an electric dipole layer develops at the
interface, having a thickness of several atomic monolayers. Macroscopically,
the spatial extension of the dipole layer is zero, but the associated dipole
392 Chapter 3. Electronic structure of semiconductor crystals with perturbations

moment has a non-zero macroscopic magnitude. As is well-known from elec-


trostatics, the potential cp exhibits a discontinuity in passing through such
a dipole layer. We denote the left boundary potential value by cp1, and the
right one by 9 2 . Then the actual positions of the valence band edges E,1
and E,2 at the heterojunction are given by the expressions

(3.232)

The difference AE, zz [E,1 - E,2] between the two valence band edges
is called valence band discontinuity or valence band offset of the material
l/material 2 heterojunction. By means of equation (3.232) the valence
band discontinuity AE, is expressed as

(3.233)

Beside the pure bulk contribution [E$ - E&], the band discontinuity also
contains the dipole contribution -e[cpi-pg]. The latter generally depends on
the properties of the interface. If one were to ignore this dipole contribution,
then AE, would obey the so-called transitivity rule which states that the
AE,-values for a succession of two heterojunctions 1/2 and 213 should
add up to the AE,-value of the combination 1/3. In practice this rule is
rarely fulfilled, which points up the importance of the dipole contribution to
A E,.
The dipole contribution -e[cpl - cpz] may be estimated by means of the
tight binding method developed in section 2.6. The bonding energy level
b considered there forms a rough measure for the average valence band
energy. In formula (2.315) b is given for a zincblende type semiconductor.
The corresponding bonding orbitals IbtR) of equation (2.316) describes the
charge transfer from the cation c (there denoted by an upper index 1)to the
anion a (there denoted by an upper index 2) in terms of the hybrid energy
difference EL - 2. If the hybrid energies were equal, no charge transfer
would occur. A semiconductor material i composed of cations ci and anions
ai, forms a big molecule of the average hybrid energy 2 +
; = (2 &/2.
Thus the charge transfer between material 1 and material 2 is governed by
the average hybrid energy difference S i - 2;. If this difference is non-zero,
valence electron charge will be transferred into the material with the lower
value of 2;. As pointed out above, the charge transfer will result in an
electrostatic potential difference at the interface. The potentials on either
side of the interface will add to the average hybrid energies and diminish their
difference. Charge will flow until the average hybrid energy difference has
been equilibrated by the potential jump. This is described by the relation

-1
ch - ecpi = F; - ecp2, (3.234)
3.7. Semiconductor microstructures 393

i.e. the average hybrid energy levels on the two sides of the heterojunction
align, and the potential contribution -e[cpl- 1,721 to the valence band offset
matches IS; - T i ] .
At this point it is advisable to recall a result concerning deep levels of
transition metal (TM) atoms obtained in section 3.5. There, it was shown
that the difference between the position E+M of a deep level of a particular
TM atom in a semiconductor material 2, and the position E&M of the same
deep level in a semiconductor material 1, equals the average hybrid energy
difference between the two materials,
ETM
2 - E&M = 7; - Sk. (3.235)
Taking the deep level positions to be measured with respect to the valence
band edges E& of the two infinite bulk materials, we denote these positions
by E ) ~ , i = 1 , 2 , so that

E ) ~ E k M - EVi.
= 0 (3.236)
Considering relations (3.234) to (3.236), the difference [ c $ ~- t+M] between
the positions of the deep level in the two materials is just the valence band
discontinuity A E , of the two materials, hence
2
AEIJ= [ETM - 61 ~ ~ 1 . (3.237)
This relation reduces the experimental determination of valence band discon-
tinuities to measurements of deep level positions of TM atoms with respect to
the valence band edge. Other experimental methods rely on measurements
of photoemission, optical or transport properties of heterojunctions. Ex-
perimental values for valence band discontinuities of several heterojunctions
are shown in Table 3.10. Often, different values are obtained by different
methods or even by different authors using the same method. This points
up the experimental difficulties in determining valence band discontinuities,
and also to the dependence of the discontinuities on the preparation of the
heterojunctions. In order to obtain theoretical values for valence band dis-
continuities, electronic structure calculations are required with an accuracy
of less than 0.1 eV. Such an accuracy is difficult to achieve so that, in many
cases, calculated valence band discontinuities have also considerable uncer-
tainty. To this day, a great deal of effort is devoted to the task of obtaining
more reIiable experimental and theoretical data for AE,, even in the case of
the most thoroughly explored heterojunction, that being between GaAs and
(Ga,Al)As alloys.

O t h e r band discontinuities
Knowing the lineup of the valence band edge at I? for a particular hetero-
junction, the lineups of all other energy levels at I' and off I' can be deter-
mined from the known bulk band structures of the two materials. Below we
394 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Junction Ge/GaAs InAs/GaSb GaAs/ZnSe HgTe/CdTe InP/CdS


AE,(eV) 0.54 -0.46 0.96 0.02 1.63
AE,(eV) 0.23 0.81 0.29 1.41 -0.39
AEg(eV) 0.77 0.35 1.25 1.43 1.24

demonstrate this taking the conduction band edges of a GaAs/Gal-,Al,As


heterojunction as an example. In Figure 3.57 the valence and conduction
band edges of the Gal-,Al,As alloy are plotted as a function of x. While
GaAs and Gal-,Al,As with x < 0.42 are direct semiconductors with both
the valence band maximum and the conduction band minimum at I, alloys
with z > 0.42 and pure AIAs form indirect materials with the valence band
maximum still at r, but the conduction band minimum at the B Z bound-
ary point X . These two cases are also to be distinguished in aligning the
conduction band edges of the heterojunction.
In the f i s t case (x < 0.42, the conduction band edges Ec1,2 of the two
materials are obtained from the relation

Ec1 = E v 1 + E ~ I , Ec2 = E v 2 + Eg2, (3.238)

where E,l,2 are the gap energies of the two materials. For the conduction
band discontinuity AE, = E,2 - E,1 of the material l/material 2 hetero-
junction, it follows that

AE, Ec2 - Ecl = -AEv + AEg, (3.239)

where AE, = E,2 - E,1 is the discontinuity of the energy gap. If AE,
and AE, are known, AE, can be calculated by means of relation (3.239).
The gap discontinuity AE, follows from the gap energies Egl,Eg2 of the two
infinite bulk materials. The lineup thus obtained for the conduction band
edge of a GaAs/Gal_,Al,As heterojunction with x < 0.42 is schematically
plotted in Figure 3.58a, together with the lineup of the valence band edge.
Often, the two band offsets AE, and AEc are expressed in percent of the
gap discontinuity AE, by means of the ratios Qv = AE,/AE, and Q c =
+
AE,/AE,. Because of AE, A E , = AE,, one has Q , Qc = 1. +
In the second case, i.e. for GaAs/AlAs or GaAs/Gal-,Al,As hetero-
junctions with alloys having an indirect gap (z > 0.42), the conduction band
3.7. Semiconductor microstructures 395

Figure 3.57: Valence band edge at


r together with the lowest conduction f
c
2.0
band levels at I? and X for Gal-xA1xAs 3
c

alloys of varying composition x. The 1.6


+
Fe(1 /2+) acceptor level is used as
energy origin following the discussion in
the main text. ( A f t e r Langer and Hein-
rich, 1985.)

0.4
-.+ ---.- Fe(1+/2+)

Composition x -

Ec2 X
r I tc2
r

I Ev2
Ev2

Figure 3.58: Conduction and valence band lineups for a GaAs/Gal-,Al,As hetero-
junction. In part (a) ( x < 0.42) the alloy gap is direct, and in part (b) (x > 0.4!2),
the alloy gap is indirect. In the latter case, the r and X levels forming the con-
duction band edges in, respectively, GaAs and Gal-,AI,As are also plotted in the
respective other material where they do not form the conduction band edge.
396 Chapter 3. Electronic structure of semiconductor crystals with perturbations

edges of the two materials occur at different k-points, that of GaAs at r,


and that of the Gal-,Al,As alloy at X . Recalling the discussion about the
stationary electron states of heterojunctions at the outset of this section, it
becomes evident that it is no longer meaningful to plot the conduction band
edge throughout the whole heterojunction as before. What may be plotted
is the lineup of the r and X levels forming the band edges in one of the
two materials. This is done in Figure 3.58b. The illustration indicates that
for the conduction band states of GaAs with wavevector r, the Gal-,Al,As
alloy region forms a barrier, and for the conduction band states of the alloy
at X , the GaAs region does so.

Types of heterojunctions

Consider two direct gap semiconductors with the valence and conduction
band edge at I?. There are several qualitatively different possibilities for the
lineup of the two band edges E,i and Eci, i = 1 , 2 (see Figure 3.59). The
conduction band edge of one material, say of materiai 1, may lie below the
conduction band edge of material 2, while simultaneously the valence band
edge of material 1 may lie above that of material 2. This case is referred
to as heterojunction of type I. In this case the gap of material 1 is located
entirely within the gap of material 2. If both the conduction and valence
band edges of a particular material, say, again 1, are, respectively, below the
two edges of material 2, one has a heterojunction of type II. The staggered
type II heterojunction occurs when we have Ec2 < E,1, and also E,1 < Ec2,
E,1 < Eva holds, and the misaligned case applies if Ec2 > E,1. In the
staggered case the heterojunction still has a gap, while in the misaligned
case the conduction band of material 1 overlaps with the valence band of
material 2 so that the gap of the heterojunction disappears. Sometimes, type
ZZI heterojunctions are defined. These are heterojunctions of type I with zero
energy gap in material 1. A look at Table 3.10 shows that the heterojunction
Ge/GaAs is of type I, CdS/InP of type 11- staggered, InAs/GaSb of type
11-misaligned, and HgTe/CdTe is of type 111.

3.7.2 Microstructures: Fabrication, classifications, examples


Heterojunctions cannot be fabricated by simply putting together two sepa-
rately made semiconductor samples with plane surfaces. If one would pro-
ceed in such a way the result would be a completely rough, polluted interface,
incapable of hosting electron and hole states which extend upon both mate-
rials, a property which has been assumed above and which turns out to be
crucial for the electronic behavior of a heterojunction. In actual practice,
the second material must be placed on top of a crystal made from the first by
continuing the crystal growth, a process referred to as epitaxial growth. In
3.7. Semiconductor microstructures 397

Type
staaaered
>r
+ EC2

P
W
C
W

Ev2 tvi

misaligned

F
x
A

d
15

0 2 0 2
material 1 material 2 material 1 material 2

Figure 3.59: Heterojunctions of type I, I1 and 111.

this kind of growth the underlying crystal, the so-called substrate, imposes
its structure onto the growing layer, in contrast to ordinary deposition of an
evaporated material which commonly results in a non-regularly structured
layer. For epitaxial growth to be possible, the two materials must have sim-
ilar crystallographic structures, and their lattice constants must be close to
each other.

Epitaxial growth

The fabrication of heterojunctions by means of epitaxy suggests to proceed


from simple structures consisting only of the substrate and the epitaxial
layer, to more complex structures by growing a second epitaxial layer of an-
other material on top of the first, a third layer on top of the second etc. If
one does so, one obtains double and multiple heterostructares. The term su-
perlattice is used if alternating layers of two materials are grown with equal
thicknesses for layers of the same material (see Figure 3.60). Two epitaxial
growth techniques are particularly important in this context, namely Molec-
ular Beam Epitaxy (MBE) and Metal Organic Vapor Deposition (MOCVD).
398 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.60: Heterostructures as grown by epitaxy: single heterostructure (a),


double heterostructure (b), multiple heterostructure (c), superlattice (d).

The principle of MBE is shown in Figure 3.61. The whole growth process
takes place in a vacuum chamber under UHV conditions (typically about
2 x 10-l' T o w ) . In this chamber, the substrate is placed in front of a
number of effusion cells which contain the various chemical elements to be
deposited on the substrate. If one, for example, wants to grow a AlAs layer
on top of a GaAs substrate, two cells are required, one for A1 and one for As,
and their shutters must be open at the same time. Doping of the layer may
be achieved by opening a third effusion cell containing the dopant atoms. If
one wants to proceed with a GaAs layer on top of the AlAs layer, one needs a
fourth cell with Ga. In order to grow a GaAs/AlAs superlattice, the shutters
of the Ga and A1 cells have to be opened and closed alternatively while the
As shutter has to be kept open all the time. There are many parameters
which influence the growth process, in particular, the temperature of the
substrate, the flux from the effusion cells, and the partial pressures of the
various elements involved. The atoms or molecules from the effusion cells
may have different sticking coefficients on the substrate. Arsenic atoms, for
example, stick much less well on GaAs then Ga atoms. Thus the As partial
pressure must be much larger than the Ga partial pressure in the epitaxy of
GaAs on a GaAs substrate.
The fact that MBE proceeds entirely in UHV may be exploited in con-
trolling the growth process. One may employ characterization techniques
which require UHV, such as RHEED for studying surface perfection, or
mass spectroscopy for analyzing the composition of the residual gas in the
growth chamber (for more on MBE see, e.g., Herman and Sitter, 1989).
Unlike MBE, MOCVD takes place in a chemical reactor at atmospheric
3.7. Semiconductor microstructures 399

RHEED GUN
LIQUID NITROGEN
COOLED SHROUDS / MAIN SHUTTER

FLUORESCENT TO VARIABLE
SCREEN SPEED MOTOR
ANDSUBSTRATE
HEATER SUPPLY

Figure 3.61: Principle of MBE growth.[After Cho and Cheng, 1981.)

pressure or slightly below. The atoms to be deposited, say A1 and As,


are provided by metal-organic gases, Al(CH3)3 (trimethyl aluminum) and
AsH3 (Arsin) in this case. The A1 and As atoms are liberated from their
compounds by means of a pyrolytic process which takes place on top of the
heated substrate. After this the atoms are chemically bound to the substrate.
In order to proceed with a layer of different chemical composition, say GaAs,
Al(CH3)3 has to be replaced by Ga(CH3)3 (trimethyl gallium) in the growth
reactor. MOCVD, unlike MBE, operates close to equilibrium conditions.
The growth velocity in MOCVD is typically somewhat larger than in MBE
because more atoms are provided to the growing layer in MOGVD than in
MBE. This makes MOCVD particularly suitable for producing devices based
on heterostructures. MBE is more universal and more responsive to control
than MOCVD. The advantages of both methods are combined in MOMBE
(Metal organic MBE): the growth process takes place in an UHV chamber,
as in MBE, but the atoms to be deposited are provided by the pyrolysis of
400 Chapter 3. Electronic structure of semiconductor crystals with perturbations

metal-organic compounds, as in MOCVD.


By means of these and some other growth techniques it became possible to
grow high-quality heterostructures of many elemental and compound semi-
conductors, as well as from semiconducting alloys. The interfaces of these
structures can be made almost abrupt, meaning that the transition from
one material to the other occurs essentially within one atomic layer. Un-
wanted impurities and structural defects can be excluded to a large extent,
both at the interface as well as in the bulk. One can grow layers up to
several millimeter thicknesses and beyond but, what is more important from
the physical point of view, one can also grow very thin layers, down to the
ultimate limit of one atomic monolayer.
It turns out that the electronic structures of multiple heterostructures
differ from those of the constituent materials if the layer thicknesses reach
the nanometer range. In subsection 3.7.4 we will prove this rigorously and
describe the modified electronic structures in greater detail. Here, we will
start with a qualitative discussion. We consider a double heterostructure
formed by a GaAs layer embedded between two Gal_,Al,As layers with
x < 0.42 (see Figure 3.62a). The electron states of this heterostructure with
energies close to the conduction band bottom will be spatially confined to
the GaAs layer. The latter is called a quantum well (QW) in this context,
and the alloy layers are referred to as barriers (for a systematic introduction
of these concepts see subsection 3.7.4). The confinement of electron states
in the quantum well raises their energy and creates discrete levels, just like
for a particle in a potential box. The same statement applies to holes in
the case of the double heterostructure of Figure 3.62a. If, instead of type
I GaAs/Gal-,Al,As double heterostructures, we consider those of type I1
(see Figure 3.62b,c), electrons and holes are no longer confined to layers of
the same material, but to layers of different materials. The central layer
in Figure 3.62b which forms a well for electrons is a barrier for holes, and
the two outermost layers which are barriers for electrons are (semi-infinite)
wells for holes. In the misaligned case of Figure 3.62c, an energy region exists
where the stationary states of the double heterostructure are mixtures of the
electron states of the two outermost layers and the hole states of the central
layer.
Another example of multiple heterostructures with modified electronic
properties is provided by superlattices composed of alternating (and suffi-
ciently thin) layers of GaAs and Gal-,Al,As with z < 0.42 (see Figure
3.62d). In this case electrons from a GaAs well layer may tunnel through
the neighboring Gal_,Al,As barrier layer reaching the next GaAs well, from
which they may tunnel through the following Gal_,Al,As barrier layer, and
so on. This leads to the formation of Bloch states and an additional en-
ergy band structure superposed upon the bulk band structure. It is called
a miniband structure. Qualitatively, the same behavior is expected for su-
Semiconductor microstructures 40 1

a) d)

Type II-staggered

Type II - misaligned
P
Q,

c b
2 2

Figure 3.62: Double heterostructure of type I acting as a quantum well both for
electrons and holes (a), and of type I1 acting as a quantum well for electrons and
a barrier for holes (b, c). In the misaligned case of type I1 (c) an energy region
occurs where electrons and holes may coexist. Parts (d), (e) and (f) of the figure
show the correponding superlattices giving rise to minibands on top of the bulk
band structure. Further discussion is given in the main text.

perlattices of the staggered type I1 shown in Figure 3.62e; solely the gap
becomes indirect in coordinate space in this case. For misaligned type I1
superlattices shown in part f of Figure 3.62, Bloch states are formed from
the electron and hole states of the respective wells.
It is not surprising that the layer thicknesses must lie in the nanometer
range for the confinement and tunneling effects discussed above to occur:
1 nm = 10 A is close to the distance between nearest neighbor atoms in a
natural crystal (in GaAs the nearest neighbor distance is about 2.5 A). Het-
erostructures with such thin layers possess, so to speak, an artificial atomic
superstructure which is likely to result in a modified electronic structure.
402 Chapter 3. Electronic structure of semiconductor crystals with perturbations

The term artificial semiconductor microstructures or just semiconductor m i -


crostructures is therefore used in this context. The term nanostructures is
also common. Besides epitaxially grown planar heterostructures and su-
perlattices having a 1-dimensional artificial microstructure (in the growth
direction), systems with a 2- or 3-dimensional microstructure are also in-
vestigated. The number of non-microstructured dimensions determines the
number of spatial degrees of freedom of electrons and holes - these systems
are termed quasi %, 1-, and 0-dimensionaL Quasi 1-dimensional systems
are also referred to as quantum wires, and quasi 0-dimensional as quan-
t u m dots. Such systems may be fabricated by means of an additional lat-
eral structuring of epitaxially grown thin layers. The more appealing nut-
ural growth of quantum wires and dots forms an area of active research at
present. One refers to this as self-organized growth (Leonard, Krishnamurthy,
Reaves, Denbaars, and Petroff (1993); Christen and Bimberg (1990); Notzel,
Ledentsov, Daweritz, Hohenstein, and Ploog (1991); Zrenner, Butov, H a p ,
Abstreiter, Bom, and Weiman (1994); Stutzmann (1995)). In this book we
concentrate on planar microstructures.
The modified electronic structure of semiconductor microstructures r e
sults in transport and optical properties which differ from those of the con-
stituent bulk materials. Moreover, these properties may be tuned to a cer-
tain extent by varying the layer thicknesses and chemical compositions of the
materials. The possibility of tailoring their properties makes semiconductor
microstructures extremely interesting subjects for micro- and optoelectron-
ics. Devices may be created with performance data superior to those of
conventional electronic components, or with functions not accessible at all
to elements made of bulk materials. High Electron Mobility Transistors
( H E M T s ) and Quantum Well (QW) laser diodes are tangible examples of
this that already exist. The concept of semiconductor microstructures was
first introduced by Esaki and Tsu (1970). Today, investigations of artificial
microstructures are the most active area of semiconductor physics.
Before we examine the electronic structure of these systems in more de-
tail, we will present a short overview of the various kinds of planar semi-
conductor microstructures. Besides heterostructure systems discussed ex-
clusively hitherto, doping microstructures will also be covered. In these, the
spatial variation of material composition is replaced by the spatial variation
of doping. We start with microstructures composed of different materials,
which are by far the most important ones.

Compositional microstructures

As a guide to finding material combinations from which compositional mi-


crostructures with modified electronic structure may be formed, one can use
a diagram which plots the energy gaps of the various semiconductors against
3.7. Semiconductor microstructures 403

3.5
' ZnS ' I \ I I I I I I I

MnSe _ .-
3.0
h

2
v 2.5
g0
2.0
>r 1.5
P
2
w
1.0

0.5
0.0
5.4 5.6 5.8 6.0 6.2 6.4 6.6
Lattice constant ( LI
Figure 3.63: Energy gap versus lattice constant for diamond and zincblende type
semiconductors. Full lines represent alloys with direct gaps, and dashed lines rep-
resent alloys with indirect gaps.

their lattice constants. Such a diagram is shown in Figure 3.63 for elemental
and compound semiconductors of diamond and zincblende structure. Lines
connecting two different materials indicate the gap energies of alloys made of
these materials. The composition of an alloy is related to its lattice constant
by means of Vegard's rule which linearly interpolates between the lattice con-
stants of the two alloy components. For two direct gap materials 1 and 2, the
gap difference AE, provides some hints about their band edge discontinu-
ities. If the gap discontinuity AE, vanishes, then band edge discontinuities
may, but need not, occur. If a gap discontinuity does exist, then at least one
of the two band edges must also exhibit a discontinuity. In general, both the
valence and conduction band discontinuities AE, and AE, differ from zero
and their magnitudes reflect the gap discontinuity to a certain extent.

Lattice mismatch

As the materials shown in Figure 3.63 are all of zincblende type, their main
404 Chapter 3. Electronic structure of semiconductor crystals with perturbations

structural differences lie in their different cubic lattice constants a. Ideally, in


epitaxial growth, the lattice constant a1 of the substrate (material 1)should
be equal to the lattice constant a2 of the layer (material 2). If they differ,
the layer material will not grow with exactly the same lattice parameters as
would a free standing bulk crystal of this material, but with a lattice constant
parallel to the layers equal to that of the substrate. One says that the layer
grows in a pseudomorphic phase. The adjustment of the parallel lattice
constant creates a mechanical strain in the layer. The corresponding strain
energy increases the total energy of the layer, and this increase grows larger
as the layer becomes thicker. Past a critical thickness d,, the accommodation
of the lattice mismatch by strain becomes energetically more costly than the
formation of lattice defects, in particular of dislocations. The strain will then
be released by the formation of dislocation lines (see section 3.2). As long
as one stays below d,, only a few dislocations occur and the layer will have
good structural perfection despite the strain. One speaks of strained layers.
The critical thickness depends on the magnitude of the lattice mismatch.
Suppose that a layer of a cubic material 2 is growing on a (100) surface of a
layer of cubic material 1, and that the cubic lattice constants a l , a:! of the
two materials are different. The relative deviation f = (a:! - a l ) / a l of these
constants is termed lattice misfit and measured in percent. For a lattice
misfit f of a few percent the critical thickness may reach values not much
smaller than 100 A. If f is less than a few tenths of a percent, the lattice
mismatch has only little effect and, in many cases, it can be completely
neglected. One speaks of lattice matched heterostructures. Otherwise one
has lattice mismatched heterostructures.
In order to grow lattice mismatched heterostructures it is important to
know the distribution of strain between layers of different thicknesses. Such
layers may be the substrate and the epitaxial layer, but also a buffer layer
on top of the substrate which is used to accommodate part of the lattice
mismatch strain. We suppose again a layer of cubic material 2 which is
growing on a (100) surface of a layer of cubic material 1. The non-vanishing
components of the stress tensors ~ ( i of) the two layers i = 1 , 2 are a,(i)
and ayy(i)with aee(i) = ayy(i),and the non-vanishing strain components
are e x s ( i ) , eyY(i) and e z z ( i ) with e r X ( i ) = eyY(i). The independent stress-
strain relations for a particular layer read

2 4 4 = cll(i)zz(i) + c12(i)[err(i) + yy(i)], (3.240)

with cll(i) and clz(i) being the elastic stiffness constants of the cubic layer
of material i. Because m t z ( i ) = 0, it follows from (3.240) that
3.7. Semiconductor microstructures 405

(3.241)

For the parallel strain components e z Z ( i ) and e y 3 / ( i ) one has

4 2 ) - &(1) = tyy(2)- E v y ( l ) = t, t = . - ai
-a2 (3.242)
ai
The total strain E defined in equation (3.242) equals the lattice misfit f
divided by 100%. Using the stress and strain components, the total elastic
energy Eda of the two layers may be calculated, with the result

(3.243)

where

ci = [Cll(i) + (1 - Ki)Cl2(i)l. (3.244)


Minimizing Eela with respect to txZ(2)(or ~ ~ ~ yields
( 1 )the
) conditions

(3.245)

According to these relations, the thin layer is more heavily strained than
the thick one, provided the elastic constants of the two layer materials are
comparable. This means, in particular, that the substrate will be almost
unstrained and the epitaxial layer will accommodate almost the whole misfit
strain. This was anticipated in the discussion above.
Another general point to be mentioned in the context of lattice mismatched
heterostructures concerns the effect of strain on the band structure. From
theoretical considerations and experimental studies it is well known that such
effects can be quite large (Bir, Pikus, 1974). To give an example, we consider
a Si layer on top of a Ge substrate. The strain components in the Si layer
are e X x = cyy = 0.04, corresponding to a lattice misfit of 4% (see below), and
e Z z = -0.03, using the elastic stiffness constants c11 = 16.1 x 1011 dyn/crn2
and c12 = 6.4 x 1011 dynlcm of Si. The degenerate heavy-light hole valence
band maximum of Si splits by 0.31 el/ under this strain. To produce the
same splitting by means of uniaxial strain applied from outside, a pressure of
73 Kbar would be necessary. This example shows that considerable changes
of the band structure are to be expected because of the lattice mismatch
strain in heterostructures. In this context, energy levels in different materials
or at different points of the first B Z can shift in dserent ways. One may take
advantage of this to adjust the band discontinuities of a heterostructure to
specified conditions. The strain becomes, so to speak, an additional degree
of freedom for tailoring the properties of a heterostructure.
406 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Particular compositional microstructures

GaAs/(Ga,Al)As

GaAs/Gal-,Al,As multiple heterostructures are the prototypes for lattice


matched type I microstructures. The valence and conduction band edges
of the Gal-,Al,As alloys have already been discussed in the context of
Figure 3.57. Here, we add some quantitative estimates. The valence band
discontinuity A E v of the GaAs/Gal-,Al,As heterojunction scales almost
linearly with x according to the relation AE, = 0.45 x x eV. For the gap
discontinuity A E , we have A E , = 1.25 x x e V , as long as we consider the
direct region x < 0.42. This gives A E , = 0.85 x x eV for the conduction band
discontinuity. With this, the valence band discontinuity takes the constant
ratio Q v = (45/125)% = 36% of the gap discontinuity, and the conduction
band discontinuity has the constant ratio Q, = 64%.
Beside providing model systems for basic research, GaAs/Gal-,Al,As
microstructures are used in devices like HEMTs, heterojunction bipolar tran-
sistors (HBPT), and QW laser diodes.

(In,Ga)As/Ga( Sb,As)

The Inl-,Ga,As/GaSbl-yAsy material system forms heterostructures of


type 11. The two band edges are lower in Inl-,Ga,As than in GaSbl-,Asy
for all compositions x and y. For small values of x and y, especially for
InAs/GaSb with x = y = 0, one has a misaligned type I1 heterostructure.
The conduction band edge of InAs is 0.14eV below the valence band edge
of GaSb. Superlattices based on InAsfGaSb heterostructures have been
the subject of intensive basic research. Under certain conditions, these SLs
exhibit metallic behavior. For Inl-,Ga,As/GaSbl-yAsY heterostructures
with larger x and y, including the trivial GaAs/GaAs homostructure, one
has staggered type I1 heterostructures, for which the conduction band edges
are above the valence band edges in both materials. Perfect lattice matching
+
is achieved if y = 0 . 9 1 8 ~ 0.082.

Si and (Si,Ge)

The lattice constants of Si and Ge are, respectively, 5.43 A and 5.65 A. This
gives the lattice mismatch of 4 %, which has already been used above. The
critical thickness d, amounts to about 30 A which is rather small. Therefore,
one also considers heterostructures between Si as substrate and Sil-,Ge,
alloys with low Ge content (x < 0.5) as epitaxial layers. In this case d, can
be 100 A or larger depending on the actual value of x. (Si, Ge) alloys may
also be used as substrates. Consider, for example, a Sil-,Gey substrate on
which a Sil-,Ge, alloy layer is grown first, followed by a pure Si layer. Both
layers are strained in this case. For 0 < y < x, the strain in the alloy layer
3.7. Semiconductor microstructures 407

is compressive, and that in the Si layer tensile. If the same Si/Sil-,Ge,


double layer is grown on Si substrate, the whole mismatch strain occurs in
the alloy layer while the Si layer remains unstrained. Because of the different
strain distributions, the band alignment of a Si/Sio.~Ge0.5heterostructure on
top of a Sio.75Geo.25 substrate differs from the band alignment of the same
Si/Si0.5Ge0.5heterostructure on top of a Si substrate. The latter is of type
I, with both the highest valence band edge and the lowest conduction band
edge in the alloy. The former still has its highest valence band edge in
the alloy, but the conduction band edge of Si is shifted down below the
conduction band edge of the alloy by the tensile strain in the Si layer. Thus
the Si/Si0.5Ge0.5 heterostructure on Si0.75Ge0.25 substrate is of type 11. The
type I1 Si/Sil-,Ge, heterostructures are better suited for applications in
electronic devices (in particular HEMTs), because here the free electrons
are hosted by the pure Si layer, where the mobility is much higher than in
the alloy layer.

(Ga,In)(As,P)/InP

Quaternary alloys of composition Inl--rGa,AsyP1-y may be thought to be


formed of the four binary compounds I d s , InP, GaAs and Gap, with com-
position ratios given, respectively, by (1-z)y, (1-z)(l-y),zy, and z(1-9).
Generalizing Vegards rule, the lattice constant of the alloy may be estimated
+ + +
as 6.058(1- z)y 5.869(1- z)(1- y) 5 . 6 5 3 ~ ~5.451~(1- y). It matches
with that of InP if the relation z = 0.189y/(0.418 - 0.013~)holds. Varying
y between 0 and 1, the lattice matched alloy transforms from pure InP to
an Inl-,Ga,As alloy with I = 0.47. Lattice matched Ino,53Gao,47As/InP
microstructures may be used to fabricate HEMTs. The fundamental en-
ergy gap of the lattice matched Inl-,Ga,As,P1-, alloy varies almost lin-
early between 1.35 eV for InP and 0.85 eV for Ino,53Gao,47As, therefore,
covering light emission wavelengths down to the technologically important
near infrared region. Using Ino.53Gao.47As/InP based microstructures, laser
diodes and photodetectors may be fabricated for optical fiber communica-
tion at 1.55 p m wavelength, which is the wavelength with minimum losses
in quartz-based fibers.

(Zn, Cd) (Se,S) /GaAs

ZnSe and GaAs form an almost unstrained type I heterostructure. If a


Znl-,Cd, S,Sel-, alloy of a certain composition z , y is combined with
a Znl-,tCd+! SY!Sel-,~ alloy of another composition x,y, one obtains a
strained heterostructure of type I. Blue-green laser diodes have been fabri-
cated from such structures, more strictly, from Zn(S,Se)/(Zn,Cd)Se/Zn (S,
Se) double heterostructures, embedded between other 11-VI-compound lay-
ers, deposited on top of a GaAs substrate.
408 Chapter 3. Electronic structure of semiconductor crystab with perturbations

(Hg,Cd)Te/CdTe

HgTe and CdTe are lattice matched because the lattice misfit is smaller than
0.4 %. Being a zero gap material, HgTe forms a type I11 heterostructure with
CdTe. The valence band offset turns out to be rather small ( M 0.02 e v ) . In
Hgl-,Cd,Te alloys the gap opens at 2 = 0.15. For 0.15 < 2 < 1 the alloys
are direct gap materials with the valence and conduction band edges at.'I
In this range, Hgl-,Cd,Te/CdTe forms type I heterostructures which, for
sufficiently small 2 , can be applied in infrared detectors.

(Pb,Sn)Te/PbTe

The lattice misfit between PbTe and SnTe is about 2 %, so that strain effects
in Pbl-,Sn,Te/PbTe heterostructures are not negligible. Pbl-,Sn,Te alloys
are direct gap materials with the conduction and valence band edges on the
first B Z boundary at L. For PbTe, the L$ state forms the valence band
edge, and the Lg-state forms the conduction band edge (see Figure 2.35).
For SnTe, this level ordering is inverted. Thus, the gap of Pbl-,Sn,Te alloys
must go through zero for some composition TO. At 77 K , one has zo M 0.4.
Thus, Pbl-, Sn,Te/PbTe heterostructures are of type I for z < 20,of type
I11 for z = zo, and again of type I for 2 > 2 0 , however, with inverted band
edges of the well material in this case. Depending on strain, type I1 situations
also seem to be possible. Microstructures based on Pbl-, Sn,Te/PbTe have
a potential for applications in infrared laser diodes.

Doping microstructures

nipi-structures

Semiconductor samples with alternating n- and p-type doped layers may ex-
hibit properties similar to those of superlattices composed of two different
materials (Esaki, Tsu, 1970; Dohler, 1972). Such doping superlattices may
be thought of as periodic arrays of pn- and np-junctions with alternating neg-
atively and positively charged intrinsic (i) regions between the neutral n- and
p-layers. In this context they are sometimes referred to as nipi - structures.
The oscillating space charge distribution gives rise to an oscillating electro-
static potential which modulates the valence and conduction band edges as
shown in Figure 3.64a. If the modulation period approaches the nanometer
range, minibands arise just as in the compositional superlattices considered
above. Doping superlattices are principally of type I1 because the two band
edges, at a given position, are shifted by the same amount of energy, namely
the electrostatic energy of an electron. Their periods, generally, cannot be
made smaller than 10 nm because of the unavoidable diffusion of dopant
atoms. The most common nipi-structures are those based on GaAs.
3.7. Semiconductor microstructures 409

a) m
+ - + - + - +
b)
+
+
+
+
+
+
+
+
+
+
+
+
+
+

-1 I
z
- - I
I
v
Z

Figure 3.64: Valence and conduction band lineup for doping microstructures in a
direct gap material for a) a nipi-structure and b) a n-type &doping structure.

&doping structures

By means of MBE, spikes of dopant atoms of only a few nanometer width


may be constructed in an otherwise undoped sample. One speaks of planar
doping or 6-doping (see, e.g., Schubert, 1994). Like nipi-structures, 6-doping
structures are always of type 11. The sheet of ionized dopant atoms forms an
electrostatic potential well which binds the emitted free (majority) carriers,
i.e. electrons in the case of n-type &doping (see Figure 3.64b), and holes in
the case of p-type &doping. The energies of these carriers become quantized
just as in the case of a quantum well. The potential well experienced by
the majority carriers represents a barrier for the minority carriers. In GaAs,
n-type &doping has been achieved, for example, by means of Si, and p-type
6-doping by means of Be.

3.7.3 Methods for electronic structure calculations


The theoretical methods which were developed to calculate the electronic
structure of bulk crystals in Chapter 2 and of clean surfaces in Chapter 3
are also suitable for artificial semiconductor microstructures. Below, we will
discuss this in the case of compositional superlattices (SLs).

Bulk methods

We consider SLs composed of two zincblende type materials as, e.g., GaAs
and AlAs. The interface is taken to be parallel to a (001) lattice plane. A
GaAs layer of the SL contains a number 2n of (001) lattice planes alterna-
tively occupied by Ga and As atoms, and the AlAs layer contains a number
2m of (001) lattice planes alternatively occupied by Al- and As atoms. The
notation (GaAs),/(AlAs), is used for such a SL. The primitive unit cell of
410 Chapter 3. Electronic structure of semiconductor crystah with perturbations

[IIO]
0 Ga OAs 8At

Figure 3.65: Primitive unit cell of a (GaAs)3/(AlAs)3SL.

the SL is spanned by the primitive lattice vectors of a (001) lattice plane,


+
namely, A1 = (a /2 )(e z ey),A2 = (u/2)(eZ - ey),and another vector A3.
Provided that (n+m) is an integer, A3 may be taken as A3 = (a/2)(n+m)ez.
The pertinent primitive vectors of the reciprocal lattice are then given by
27r 27r 4n
B1 = -(ez
a - ey), B2 = -(ez
a + ey), B~ =
(n +m)a ez. (3.246)

+
The volume of the first B Z of the SL is a ( n m)-th fraction of that of the
first bulk B Z . In Figures 3.65 and 3.66we show, respectively, the primitive
unit cell and the first B Z of a (GaAs)g/(AlAs)~SL as an example.
The stationary states of the SL are Bloch states with quasi-wavevectors
k of the first SL B Z . The energy eigenvalues form bands in this B Z , re-
ferred to as minibands in regard to k-dispersion parallel t o the SL axis, and
as subbunds if k-dispersion parallel to the SL layers is considered. The mini-
bands arise from bulk bands folded back upon the first SL B Z . As the bulk
+
B Z encompasses ( n m ) first SL B Z s , one bulk band gives rise to ( n m) +
SL minibands. The minibands are separated by m i n i g a p s which occur at
the Bragg reflection planes of the reciprocal SL lattice perpendicular to the
z-axis. The minigaps become wider for stronger perturbations of the bulk
crystals due to the superlattice structure, i.e. the larger the difference of the
periodic potentials of the two bulk crystals, and the shorter the SL period
is (as long as it does not become too short), the wider the minigaps open.
All methods for determining band structure of bulk crystals can also be
used to calculate the band structure of SLs, the only difference being the
larger size of the primitive unit cell. Figure 3.67 shows the results of such a
calculation performed by means of the TB method. The valence band disper-
sion of a (GaAs)a/(Ga0,~A10,3As)3SL is plotted along the r - Z symmetry
3.7. Semiconductor microstructures 41 1

1".

Figure 3.66: First B Z of the (GaAs)3/(AlAs)3 SL of Figure 3.65.

line of the first SL B Z and, to demonstrate the folding character of the SL


band structure, also along the r - X line of the first bulk B Z . Important
properties of the corresponding SL eigenfunctions are illustrated in Figure
3.68 for a (GaAs)7/(AlAs)T SL. The lowest valence and conduction band
states are represented with respect to their spatial variations and symmetry
characters. While the valence band states are well localized in the GaAs
layers and are dominantly composed of GaAs-r-states, both AlAs-X-states
and GaAs-r-states contribute to the conduction band states of the SL. This
illustrates the general discussion of subsection 3.7.1 about the existence of
eigenstates of heterojunctions which arise from bulk states with different
k-vectors in the two materials .
Bulk methods may also be used to calculate the electronic structure of
non-periodic microstructures such as GaAs/AlAs single heterostructures or
(Ga, Al)As/GaAs/(Ga, A1)As double heterostructures. In these cases one
may use a procedure similar to the slab method for surfaces: At the outset,
one defines a slab with a sufficiently thick but finite GaAs layer and forms
a single heterostructure with a sufficiently thick but finite AlAs layer, and
then repeats this slab periodically. In the case of the (Ga, Al)As/GaAs/(Ga,
A1)As double heterostructure, one embeds the GaAs layer between thick but
finite (Ga, A1)As layers, and repeats the slab thusly formed periodically. In
this way an SL is simulated whose electronic structure can be calculated by
means of any of the bulk band structure calculation methods.
However, in many cases the application of bulk methods to artificial mi-
crostructures is not appropriate. These methods yield the totality of energy
bands of a microstructure in all parts of the first B Z , while in most cases
only a small number of minibands are of interest because only these undergo
changes in comparison with the back-folded bulk bands. Moreover, even the
412 Chapter 3. Electronic structure of semiconductor crystals with perturbations

-to
3 -1
-A
ol
!i
W
-2

-3

-4

-F;

Figure 3.67 Valence band structure of a (GaAs)3/(Gao.7Alo.3As)3SL calculated


by means of the TB method, and plotted upon the r - 2 symmetry line of the first
SL B Z (left) and the I?-X line of the first bulk B Z (right). (After Riicker, Hanke,
Bechstedt, and Enderlein, 1986.)

changes of these few minibands are small. The bulk methods are often not
accurate enough to reproduce them sufficiently well. The situation is similar
to the case of the shallow levels of impurity atoms - if one would try to ob-
tain these levels by means of a full bulk band structure calculation method
the same difficulties would occur. Fortunately, another method, namely the
effective mass theory, exists in the shallow level case. It is particularly well
tailored to calculate the small changes of electronic structure occurring at
the conduction band minimum and the valence band maximum, where the
kinetic energy is small enough to allow the perturbation potential to produce
measurable effects. The situation in the case of artificial microstructures is
comparable: the perturbation potential is relatively weak, and changes of
the band structure are expected only for selected bands and in the vicinity
3.7. Semiconductor microstructures 413

GaAs A1As r X

Figure 3.68: Localization of the lowest valence and conduction band states of
the SL at the center r of the first SL B Z (left), and the symmetry character of
these states (right), for a (GaAs)7/(A1As)7SL. While the valence band states are
well localized in the GaAs layers and are dominantly composed of GaAs-r-states,
conduction band states extend over both layers, and are composed of both AlAs-
X-states and GaAs-I'-states. h l = -0.11 eV,h2 = -0.19 e V , el = 1.75 eV,e2 =
1.77 eV, e3 = 1.88 eV.(After Rucker, 1985.)

of critical points. There are, however, also differences between the shallow
level problem and the microstructure problem. The perturbation potential,
i.e. the difference of the periodic oneelectron potentials of the two m a t e
rials of a heterostructure, is far from being smooth on the atomic length
scale. Even if one averages out the microscopic potential fluctuations over
a unit cell, the average potential difference has still an abrupt change at an
interface. For this reason one has to determine at the outset whether or not
the effective mass theory developed in section 3.3 can in fact be applied to
artificial microstructures. We will address this question below (for a detailed
discussion see, e.g., Burt, 1992).
414 Chapter 3. Electronic structure of semiconductor crystab with perturbations

Effective mass theory for microstructures

Consider a single heterostructure. There is no doubt that the effective mass


theory may be applied to each of the two infinite half spaces of this structure
filled with material 1 for z < 0 and with material 2 for z > 0, provided the
conditions of validity of effective mass theory are satisfied in each of the two
regions separately, which we will assume. Accordingly, we restrict ourselves
to eigenstates of the heterojunction having energies in the vicinity of the
band edges E,1 and E,2 of the two bulk materials. Both edges should occur
at the centers I? of the respective bulk B Z s . The two band edges at
are assumed to be non-degenerate and isotropic (the degenerate case will
be treated separately). Furthermore, we suppose that, at critical points
off r, all bulk band energy levels are far removed from the energies of the
eigenstates under consideration. Then these eigenstates are composed of only
bulk states with k-vectors in the vicinity of r, just as is assumed in effective
mass theory (in fact this theory assumes composition of bulk states from the
vicinity of a particular critical point which need not necessarily be f). If the
eigenstates were composed of bulk states from different parts of the B Z , say
from f and X as happens in the case of the conduction band states of the
(GaAs)7/(AlAs)i. SL shown in Figure 3.68, then the effective mass theory
could not be applied. If, however, an eigenstate of the heterostructure may
be formed only from bulk states at an off-center critical point like X , then
the effective mass theory is applicable as well.

Non-degenerate band edges

In the case under consideration, the eigenfunctions & l ( x ) and $ 4of~ the)
two material regions 1 and 2 having energies, respectively, close to the band
edges Evl and E,2 at f,may be written as

The two envelope functions F,l(x) and Fv2(x) obey the effective mass equa-
tions

F,l(x) = E,F,i(x), x in m a t e r i a l 1, (3.248)

F V ~ ( X=) E , F , ~ ( x ) , x in m a t e r i a l 2. (3.249)

Formally, these equations may be written as one equation for the envelope
function
3.7. Semiconductor microstructures 415

(3.250)

of the whole heterostructure by introducing a z-dependent effective mass


m * ( z ) and a z-dependent band edge E,(z) defined, respectively, by

We write this one equation as

(3.252)

The true wavefunction $,(x) of the heterostructure is given by the expression

(3.253)

(3.254)

is the Bloch factor of the heterostructure. Clearly, equation (3.252) holds


only within the two material half spaces, but not at the interface. Thus it
determines the possible envelope function solutions for z < 0 and z > 0. At
z = 0, the solutions for the two half spaces must be connected in an appro-
priate way to form an envelope eigenfunction for the entire heterostructure.
To do so, matching conditions are required, which we will now discuss.
Firstly, we derive a condition for the change of the envelope function
across the interface. To this end we consider the interface behavior of the
Bloch factor ud(x) in equation (3.253) for the total wavefunction &,(x). If
this factor was continuous at the interface, the envelope function F,(x) also
would have to be continuous there since the total wavefunction @,(x) must
be continuous everywhere. Thus the condition

Fv(z,Y, 0 - 6 ) = F,(z, Y, 0 +6) (3.255)


should hold in the limit 6 --+ 0. Unfortunately, no general proof exists for the
continuity of the Bloch factor at a heterostructure interface. The only case in
which such a continuity is assured is that of a 'heterostructure' composed of
two identical materials. Having this obvious result in mind, it is commonly
416 Chapter 3. Electronic structure of semiconductor crystals with perturbations

argued that the Bloch factor should be continuous at an interface, at least


in an approximate sense, if the two materials of the heterostructure are not
too different from each other with respect to their energy band structures.
This happens, for example, in the case of GaAs/Gal-,Al,As heterostruc-
tures with sufficiently low z-values. It turns out, however, that effective
mass calculations based on envelope function continuity, also yield correct
results for heterostructures made of materials with considerably different
band structures. Thus, similarity of band structures, while being sufficient,
does not seems to be necessary for the applicability of effective mass the-
ory to heterostructures. In fact, there have been various attempts in the
literature to justify the continuity condition (3.255) without using the band
structure similarity argument. In our opinion, it is essential to realize that
the Bloch factor ul/o(x) in equation (3.252) occurs in an eigenfunction of
the entire heterostructure rather than in an eigenfunction of the two infinite
bulk crystals, as is tacitly assumed in the above discussion. For an infinite
bulk crystal, the Bloch factor is that particular solution of equation (2.136)
which obeys the lattice periodicity condition. Without demanding satisfac-
tion of this condition, there exists a variety of solutions of equation (2.136).
In a heterostructure, the two Bloch factors at the interface need not be pe-
riodic with respect to z because the lattice periodicity in the z-direction is
perturbed. This freedom may be used to satisfy the continuity condition for
the Bloch factor at the interface.
Secondly, we derive a condition for matching the interface values of the
first derivative of the envelope function with respect to t. To this end we use
the physically obvious fact that the probability current density joined with
the total wavefunction $ J ~ ( X )must be the same on the left and right of the
interface. This must also hold after averaging with respect to a primitive
unit cell, i.e. with respect to a region where the envelope function is almost
constant. Applying the well-known quantum mechanical expression for the
current density one obtains two contributions. One is due to the gradient
of the Bloch factor and it vanishes after averaging. The other contribution
arises from the gradient of the envelope function. It is multiplied by the
squared modulus of the Bloch factor, which yields 1 after averaging because
of normalization. This means that the average current density in state $v(x)
is obtained by applying the quantum mechanical expression to the envelope
function Fv(x)alone rather than to the total wavefunction &,(x). Since
the average current density must be continuous at the interface, and since
the envelope function itself has this property because of relation (3.255), we
arrive at the conclusion that the first derivative of the wavefunction must
obey the relation
3.7. Semiconductor microstructures 417

which is known as the BenDaniel-Duke boundary condition. This condit,ion


follows automatically if the Schrodinger equation (3.252) is also applied at
z = 0 after the kinetic energy term has been replaced in accordance with

h2 h2
_- v2 -.+ --v.- v. (3.25 7)
2m*(z) 2 m*(z)
With this replacement the effective mass equation (3.252) becomes

Integrating this equation with respect to z from -6 to +6 one obtains the


boundary condition (3.256).
The kinetic energy operator of the Schrodinger equation (3.257) is Hermi-
tian, as any reasonable kinetic energy operator must be in order to avoid
complex energy eigenvalues. The original form -[l/2m*(z)]V2 of the kinetic
energy operator is not Hermitian, and must be rejected. Its replacement by
the right hand side of relation (3.256) cannot be justified, however, by the
hermiticity demand alone since there are also other ways to introduce her-
miticity. In fact, one can easily demonstrate that any operator of the form
-(F,2/4)[m*a(z)Vm*~(z).Vm*~(z)+m*~(z)Vm*~(z).Vm*a(z)] is Hermitian
+ +
if a,p, y are real numbers obeying the relation (Y ,B y = -1. This implies
that the boundary condition (3.256) involves more than just the hermiticity
of the kinetic energy operator.
In the effective mass equation (3.258) the band edge plays the role of an
external potential. Since it depends only on z , and since the effective mass
does so also, the dependence of the envelope function on the component XI[
of x parallel to the layers may be taken in the form of a plane wave of parallel
wavevector kll, therefore as

(3.259)

For F u ( z ) the Schrodinger equation follows as

where EL is given by

h2
EL = E , - -k i (3.261)
2m*,
with l/E; as the average inverse effective mass (m;:' + m;T1)/2.
418 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Degenerate band edges


@
In the case of degenerate band edges E v ( z ) ,the Hamiltonians H i of the two
materials i = 1,2, are of the general form

(3.262)

derived in equation (3.73), and the components Fym(X) of the envelope func-
*
tions F~ ( x )obey the effective mass equations

[-~g$r(z)aaaat + ~ v ( z ) 6 m m iFumt(x)
] = ~ v ~ v r n ( x z) ,# 0,
m a d
(3.263)
with DE$(z) = DYZAl for z < 0, and Dg$l(z) = D;ELl for z > 0. The
solution of equation (3.263) may again be taken as a plane wave

(3.264)
=+
parallel to the layers. The boundary conditions for F v ( x ) at the interface
follow in the same way as in the non-degenerate case. The z-dependent
* *
factor F~ (2) of F,, ( x )in equation (3.264) must be continuous at z = 0, i.e.
* *
F U (z)lz=-6 = F u (z)I~=+6 (3.265)
must hold for 6 + 0. To determine the condition for the derivative with
respect to z , one defines the Hermitian Hamiltonian matrix

(3.266)

($
which exists for any value of z including z = 0 and equals H 1 for z < 0 and
e
H z for z > 0. Integrating the effective mass equation for this generalized
Hamiltonian with respect to z over a small interval across the interface leads
to the conclusion that the expressions

must be continuous at the interface z = 0.


Below, these general results are specified for the fourfold degenerate rs va-
lence band edge of diamond and zincblende type materials. Using the 4 x 4
Luttinger-Kohn Hamiltonian of equation (3.74), the effective mass equation
(3.263) becomes
3.7. Semiconductor microstructures 419

(3.268)

The quantities Q ( z ) ,f i ( z ) ,g ( z ) , ? ( z ) here are the differential operators

k(z)= --a
[y2(zj(k: - k i ) - 2i73(z)k,kg] . (3.272)

The constants M , L , N in equations (3.75) to (3.78) have been replaced by


the Luttinger parameters y1,72,y3 using equations (2.381). From equation
(3.267) one finds that the following combinations of the components F m ( z )
and their derivatives FLm(z) with respect to z must be continuous at the
interface z = 0:
420 Chapter 3. Electronic structure of semiconductor crystals with perturbations

L
XJ

Figure 3.69: Geometry of the SL whose stationary electron and holes states
are calculated in the text.

3.7.4 E l e c t r o n i c structure of particular microstructures

Compositional superIat tices and q u a n t u m wells

We consider a SL composed of two zincblende type semiconductor materials


1 and 2 with the conduction band edges Ecl and Ec2 located at r. To be
specific, we may associate 1with GaAs, and 2 with a Gal-,Al,As alloy
for z < 0.42. The thicknesses of the two material layers are denoted by d l
+
and d 2 , and the lattice constant in z-direction by d , with d = d l d 2 ( s e e
Figure 3.69). We want to calculate the stationary electron states of this SL
with energies in the vicinity of the conduction band edges Ecl and E,2 using
the effective mass theory derived above. These states are the ones which
would host the free electrons of the SL, if there were any. We suppose that
there are none, as we did before.

Electron states

Effective mass equation and its solutions

For the zincblende type materials we are considering, the conduction bands
are non-degenerate and isotropic in the vicinity of I. The two effective
electron masses m:l and mE2 generally differ, but we will initially assume
that they are equal, denoting their common value by m;. In the case of
SLs composed of GaAs and Gal_,Al,As this is a reasonable approximation.
Later, we discuss the modifications which occur if m;l and m:2 are different.
The effective mass equation for the electron envelope function Fc(x)of the
3.7. Semiconductor microstructures 421

thus specified SL follows from equation (3.252) if we set v = c and

f o r Id < z < ld+dl


Ec(z) = (3.274)
Egl +AE, f o r Id + d l < z < (1 + l ) d .
Here 1 is the integer index of the various SL unit cells, -co < 1 < 00.
The envelope function F,(x) may be taken in the form Fckl, (x)of equation
(3.260) describing a plane wave of wavevector kll parallel to the layers. The
z-dependent factor F,(z) of Fckl,(x)obeys the equation

where E; is given by

EA = E , - (3.276)

In solving this equation, the boundary conditions

Fc(ld + d l - 0 ) = F,(ld + d l + 0 ) ,
d d
-Fc(ld
dz
- 0) = -F,(ld
dz
+ 0),
d
-F,(ld
dz
+d l - (3.277)

have to be satisfied as well as the normalization condition with respect to


the periodicity interval of length C 2i = Ld ( L denotes the number of SL
unit cells in a periodicity region of volume 0). Applying the Bloch theorem
to the SL under consideration it follows that F,(z) may be written as a
lattice-periodically modulated plane wave

F,(z) = Fck(z) = Lfe ii k r U c k ( z ) , (3.278)

where k is the quasi-wavevector component parallel to z , and U c k ( z )is the su-


perlattice Bloch factor. The component k varies within the (1-dimensional)
first B Z of the superlattice between - r / d and r l d . It must have the form
( 2 a / L d )x (0, f l , f 2 , . . .) in order to guaranty the periodicity of F&) with
respect to the periodicity interval. The Bloch factor U c k ( z ) of the superlat-
tice has to be distinguished from the Bloch factor uco(x) of the two bulk
422 Chapter 3. Electronic structure of semiconductor crystals with perturbations

materials defined in equation (3.254). The occurrence of two different Bloch


factors reflects the fact that two periodic potentials are present in a super-
lattice, one due to the natural crystal structure, and one due to the artificial
superstructure. The total wavefunction &(x) &k(x) contains the product
of both. It reads

(3.279)

The effective mass equation (3.275), the matching conditions (3.277), and
the Bloch condition (3.278) define an energy eigenvalue problem which differs
from the well-known Kronig-Penney problem only by notation. The latter
problem constitutes an exercise of elementary quantum mechanics and is
commonly used to demonstrate the existence of energy bands. In the case
of SLs this exercise takes on real physical meaning. In the following we
sketch its solution. We restrict ourselves to energy values 0 < E' < A E ,
or E,1 < E < E,1 + +
AE, (7i2/2mr)ki, meaning energies below the lowest
allowed energy in material 2 for a given value of kll.
According to classical mechanics, an electron with such an energy cannot
penetrate within the superlattice layers made of material 2, it will be confined
to layers consisting of material 1. If it hits the interface to material 2, then
it will be reflected back to the interior of its material 1 layer. In quantum
mechanics, the reflection is not complete, there is a certain probability for the
electron to tunnel through a material 2 layer and reach the next neighboring
material 1layer. Although the probability for an electron to stay in material
1 is not unity, as in classical mechanics, but smaller, it is still much larger
than that for material 2. As has already been mentioned, one uses the
terms quantum wells for the layers of material 1, and bamiers for the layers
of material 2. In the wells and barriers the wavefunction F,]F(z)is given
by different expressions. For the wells 0 < z < d l , Schrodinger's equation
(3.276) yields

F,]F(z)= ale iKZ + ble-iKz, 0 < z < dl, (3.280)


with K as a real number which determines the energy E' by means of the
relation

E ,I = - h2
K2 (3.281)
2 m ~ '
and a l , bl as coefficients which are still to be determined. For the barrier
d l < z < d one has

Fck(z)= a2enz + b2e-nz, dl < z < d, (3.282)


where we have set
3.7. Semiconductor microstructures 423

2 2mE
KE.= -(AE, - EL). (3.283)
Ti2
The four coefficients a l , b l , a2, b2 are governed by a set of linear equations
which follows from the 4 matching conditions (3.277). These equations read

For this set to have a non-trivial solution, the determinant of the matrix of
exponentials must vanish. This yields the secular equation

tc2 - K~
cos(Kd1) cosh(tid2) +- 2Kn:
sin(Kd)l sinh(~d2)= cos(kd). (3.285)

Allowed K - and &-valuesmust satisfy equation (3.285). The values of K and


K are not independent of each other, both are determined by the energy E L
through equations (3.281) and (3.283), respectively. Thus equation (3.285)
represents a condition for EL. It coincides with that of the Kronig-Penney
problem. As we know from the solution of the latter, there are discrete energy
eigenvalues EL, n = 1,2,. . ., for a given quasi wavenumber k. They split into
energy bands E&(k), separated by energy gaps, if k varies within the first
B Z . To obtain the total energy eigenvalues E, = Ecn(k) of the superlattice,
according to equation (3.276), one still has to add E,1 +
(Ti2/2m*)kll to
E L ( k ) , with the result

h2
Em(k) = E&(k) + Egi + -kll
2mr
(3.286)

Unlike the Kronig-Penney problem, these energy bands exhibit an additional


dispersion with respect to the wavevector component kll along the SL layers.
Commonly, the energy bands and gaps for a fixed value of kll, but with
varying values of k , are termed minibands and minigaps. Bands which arise
for fixed IF and varying kll are termed subbands. The kll or subband dispersion
is due to the natural crystal structure, while the k or miniband dispersion
follows from the artificial microstructure. In Figure 3.70, central part, the
sub- and minibands of a SL are schematically plotted. The effective masses
of the minibands become negative at the boundaries of the superlattice B Z
in the case of odd band numbers n, and at the center of the B Z for even n.
424 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.70: Mini- and subbands of a superlattice. Also shown are the two
limiting cases of extremely wide and/or high barriers and of narrow and/or
flat barriers. In the lower part of the figure the wavefunctions are illustrated.

Multiple quantum well versus superlattice behavior

Figure 3.70 also shows two enlightening limiting cases. If the thickness d2
of the barriers between the wells becomes large, and/or the barriers AE,
are high, then the hyperbolic terms in equation (3.285) dominate over the
cos(kd)-term on the right hand side which is responsible for the k-dispersion.
Neglecting the latter, the minibands degenerate into discrete energy levels,
termed sublevels. These levels are just the discrete energy eigenvalues of a
single quantum well. Obviously, the superlattice decomposes into isolated
quantum wells in the limiting case under consideration. The electron states
are almost completely confined within a well, and practically zero within the
barriers. One uses the term multiple quantum well for a superlattice in this
limit.
For infinitely high potential walls the secular equation (3.285) becomes
sinKd1 = 0. The solutions of this equation are K-values of the general form
K = (7r/dl)nwhere n is an integer. They give rise to the energy eigenvalues
2
Elcn = , n = 1 , 2 , .. . , (3.287)
2mE
3.7. Semiconductor microstructures 425

and Bloch factors

(3.288)

The confinement to the well region is complete in this limit. Assuming


d l = 100 A and mE = 0.067 m , one gets a value of 57 meV for the first energy
level ELl. This means that the ground state of the quantum well is shifted
up by this energy amount as compared to the conduction band bottom of
the infinite bulk crystal. This shift is a consequence of the localization of an
electron within the quantum well which leads to a momentum uncertainty
because of Heisenberg's uncertainty principle.
We consider next the opposite limiting case, that of extremely narrow
and/or flat barriers (see Figure 3.70). In zero-th approximation the secular
equation (3.285) takes the form k = K . In this limit the minigaps vanish
completely, and the superstructure which distinguishes the superlattice from
a natural crystal has no effect at all. In the first non-vanishing approximation
with respect to the SL potential, gaps are opened at the Bragg reflection
points k = ( n n / d ) of the 1-dimensional reciprocal lattice of the SL. For
the energy splitting between the two bands E&+l(n/d) and E & ( n / d ) , the
perturbation theory developed in section 2.4 yields

E&+1 (:) -E L (:) 1


= -AEc
7rn sin ( y d g ) . (3.289)
According to this expression, the minigaps are proportional to the conduc-
tion band discontinuity AE,. They decrease with band number n. Thus,
for large n, there are no substantial energy gaps even if sufficiently wide and
high barriers exist. One may say that for minigaps to occur the product of
wavevector k = ( x n / d ) and superlattice period d should not be large com-
pared to 1. In other words, the superlattice period d should not be large
in comparison with the electron wavelength X = 2 r / k corresponding to k.
This proves and specifies the expectation stated at the outset of this section,
that substantial changes of electronic structure are to be expected if the
superstructure occurs on a sufficiently small length scale. The length scale
turns out to be that of the de Broglie wavelength X of an electron. For a
conduction band electron of silicon, having an energy equal to the average
thermal energy $kT at room temperature, one gets A =" 20 nm. Similar
values are obtained for other semiconductors. Thus the layer thicknesses
of superlattices have to be on the order of, or smaller than, 10 nm in or-
der that superlattice effects on free carriers may become important at room
temperature.
The limit of narrow and flat barriers is the true superlattice case, as
distinguished by the coupling of quantum wells with one another. The en-
426 Chapter 3. Electronic structure of semiconductor crystals with perturbations

ergy eigenfunctions also have non-vanishing values within the barriers, in


contrast to multiple quantum wells where these values are almost zero. In
a superlattice, an electron having a particular allowed energy, may tunnel
from its quantum well through the barrier to the neighboring well. Since
there it encounters the same allowed energy value from which it may tunnel
back, the tunneling will be amplified, as in a Fabry-Perot resonator in the
case of light. One has so called r e s o n a n t tunneling. It causes the discrete
energy levels of isolated quantum wells to broaden into bands.

Electron density of states of SLs and QWs

The general expression (2.212) for the density of states (DOS) p ( E ) of a 3D


crystal may be immediately applied to a SL. To this end one has to replace
the energy band structure E,(k) in formula (2.212) by the expression given
in equation (3.286). Then the DOS p s ~ ( Eof) the SL is obtained as

To evaluate this expression further, the miniband dispersion EA(k) has to


be known. In the particular case of isolated QWs the dispersion vanishes
completely, and the DOS of equation (3.291) takes the form of the DOS
P M Q W ( E )of a MQW structure,

(3.291)

Each subband n gives rise to a step-like partial DOS, being zero below the
+
band bottom E L E g l , and non-zero but constant above. Such behavior
of the DOS of an electron which is free to move only in two dimensions
was already found in section 2.5. The summation in expression (3.291) for
p ~ over ~ all w subbands results in a staircase-like DOS as schematically
shown in Figure 3.71. The heights of the steps have the constant value
( m * / n 2 h 2 ) ( n / d )while
, the step widths depend on n. If the barriers of
the QW are taken to be infinite, then the energy levels EAn are given by
( h . 2 / 2 m E ) ( n ~ / d 1 and
) 2 , the step widths of the staircase scale with ( n / d ~ ) ~ .
For very wide QWs, i.e. in the limit d l -+ 00, the staircase-like DOS trans-
forms into the smooth square-root-like DOS p 3 ~ ( Eof ) an electron free to
move in three dimensions. Equation (3.291) yields

(3.292)
3.7. Semiconductor microstructures 427

7 3
3
0
1

3
0 2
Q

Figure 3.71: Electron density of states PQW of a QW as function of energy.

with f = limd,,,(dl/d) being the geometrical fraction of material 1 regions


in the infinite MQW. The staircases of the DOS in Figure 3.71 are always
below or at the limiting fractional 3D DOS f p j ~ ( E ) .The case when they
are at the DOS occurs for energies E at the subband bottoms.

Refinements

In the above treatment of the electron states of a SL, the effective masses
mzl and mE2 were assumed to be the same in the two materials. If this is not
the case, the following modifications will occur. First, in the boundary con-
ditions (3.277) for the derivative ( d F , / d z ) , different mass factors will appear
on the two sides of the interface and they will not cancel. For the secular
equation (3.285) this means that all K ' s except those which are factors in
arguments of trigonometric functions, have to be replaced by (m&'mal)K.
Second, in solving the secular equation, the relations (3.281) and (3.283)
between K and E i and, respectively, n and E' must be written with m;
and mE2 instead with mE. Finally, the kll-dispersion of the total energy
eigenvalue in relation (3.286) becomes different in the two material layers.
This results in an additional contribution (h2/2)[mZ-'(z) - E ~ - ' ] k to ~ the
SL potential, as may be seen from equation (3.260). The additional term
introduces a kll-dependence of the envelope function F& which otherwise
depends only on the z-component k .
Another simplification made in the above treatment of the electron states
of a SL, was the assumption that no electrons are available to occupy the SL
428 Chapter 3. Electronic structure of semiconductor crystals with perturbations

0.5 x102cm-2 1.5 x102cm-2 6 xlO*cm-*


t

Figure 3.72: Self-consistently calculated potential profiles and energy levels for
a Alo.aGao..rAs QW of 100 A width containing different numbers of electrons per
cm-2.

bands. If there are such electrons present, the potential seen by an electron
will change due to its interaction with other electrons in the SL bands. As in
the case of bulk semiconductors, this potential change may be decomposed
into a Hartree and an exchange-correlation part. The expressions derived for
these potentials in Chapter 2 also apply here if the stationary states of the
bulk crystal are replaced by the stationary states of the SL. As the latter
states are only known after the effective mass equation has been solved,
the inclusion of the Hartree and exchange-correlation potentials must be
performed in a self-consistent way. The net effect of the two potential parts
is repulsive, thus the SL potential well will flatten. For a given number of
electrons in SL states, this flattening will be enhanced by greater localization
of the electrons in the well regions. It will be particularly pronounced for
isolated QWs. In Figure 3.72 we show self-consistently calculated potential
profiles and energy levels for a Alo.3Gao.7As QW of 100 A width, containing
different numbers ng of electrons per ~ r n - ~The. well bottom bends up with
rising n s , and the energy levels are shifted to higher energies.
Hole states
We consider a QW structure composed of zincblende type well and barrier
materials which are both described within the Luttinger-Kohn model. The
effective mass equation for the hole states of such a QW with energies close to
the well bottom follows from equation (3.268) by installing there the valence
band edge profle E,(z):

0 forO<z<d
E,(z) = (3.293)
AE, f o r z < 0 and z > d.
3.7. Semiconductor microstructures 429

Figure 3.73: Hole subbands


of a 100 A wide Alo.sGao.TAs
QW. ( A f t e r Altarelli, Eken-
berg, and Fasolino, 1985.)

In this way, a set of second order differential equations for the four hole
envelope function components arises. The solutions must be normalized
and must guarantee for the continuity of the expressions (3.273) at the two
interfaces. Solutions of this kind can be found in different ways. One can,
for example, numerically calculate normalized solutions for various values
of E and kll, valid in one of the three regions -m < 0, 0 < z < d and
z < +m. Matching these solutions in such a way that the continuity of
the expressions (3.273) is guaranteed for a given value of kll, will only be
possible for certain discrete energy values E,,(kll). These are the subband
energy eigenvalues which are sought. Subband structures calculated in this
way are shown in Figure 3.73 for a A10.3Ga0.7As QW of 100 A width. The
lowest subband ( h h l ) in the vicinity of kll = 0 arises mainly from hole states
of GaAs with the heavy hole mass parallel to z , and the second level ( I h l )
mainly from hole states with the light hole mass parallel to z. Perpendicular
to z , the curvature of the hhl subband at starts stronger than that of
the Ihl subband. Thus the two bands would cross, if this were allowed. In
reality, the bands repel each other leading to the anticrossing behavior seen
in Figure 3.73. The alternating heavy- and light-hole-like curvatures parallel
and perpendicular to z are consequences of the Luttinger-Kohn Hamiltonian
(2.374).
If the subbands of the QW are partially filled by holes, then hole-hole inter-
:::
430 Chapter 3. Electronic structure of semiconductor crystals with perturbations

p-type &well r-type &well


50

8.0x 10 cm-*
-50 -250
-200 -100 0 100 200 -200 -100 0 100 200

z (4 z (A)

Figure 3.74: Potential pofiles and energy levels of an isolated p-type &doping layer
(left hand side) and an isolated n-type-&dopinglayer (right hand side) of the same
sheet dopant concentration 8 x 10l2 ~ r n - ~The
. Fermi energy is shown by dotted
lines. (After Sipahi, Enderlein, Scolfaro, and Leite, 1996.)

action effects become important which, as in the case of electrons, require


self-consistent calculations.

&doping structures

Whereas in compositional microstructures free carriers may, but need not,


be present (they were omitted above), &doping microstructures do not exist
without such carriers because the doping layer would be neutral if none of
the doping atoms would be ionized. This implies that potential profiles and
energy levels of doping microstructures cannot be calculated without taking
carrier-carrier interaction into account. The main effect of this interaction
turns out to be the screening of the Coulomb potential created by the ionized
dopant atoms. In Figure 3.74 we show the resulting potential wells for an
isolated p-type &doping layer (left hand side) and an isolated n-type 6-
doping layer (right hand side), together with the sublevel and Fermi level
energies. The dopant atoms are taken to be completely ionized. The lower
3.7. Semiconductor microstructures 43 1

Figure 3.75: Calculated cur- 100


rent-voltage characteristics of va-
rious GaAs/ AlAs SLs with the 80
electric field parallel to the SL -
axes. Curves 1to 4 correspond to
SLs of different miniband widths
>a 60

A. (After Lea, Horing, and Cui,


1991.)
=
0

0
40
>
20

0
0 10 20 30 40 SO 60 70 I D
E [kV/cml

well depth in the p-type case as compared to the n-type case results mainly
from the fact that heavy holes are more strongly localized at the ionized
dopant sheet and thus are more effective in screening out the sheet potential
then the lighter electrons.

Macroscopic manifestations of the electron and hole states of semi-


conductor microstructures

The modifications of electron and hole states of semiconductor microstruc-


tures manifest themselves in the macroscopic electronic properties of these
structures. The more detailed treatment of these properties is beyond the
scope of this volume. Here we will give only two characteristic examples,
one concerning electric transport and another concerning optical properties.
Figure 3.75 shows the calculated drift velocity V d versus electric field char-
acteristics of GaAs/ AlAs SLs with the electric field E directed along the
SL axes. Above a critical field value at which the V d peaks to its maximum
value, the drift velocity decreases with increasing field strength, i.e, the dif-
ferential conductivity becomes negative. Such a behavior, which originally
was predicted by Esaki and Tsu (1970), has been confirmed in more rigor-
ous calculations by Lei, Horing and Cui (1991). Experimentally, a sublinear
increase of V d with E has been found above a critical field value, as shown in
Figure 3.76. If corrected for effects of the Ohmic contacts and macroscopic
electric field inhomogeneities, the experimental V d versus E characteristics
clearly reveals the negative differential drift velocity predicted by Esaki and
Tsu. The underlying physics can be understood easily.
Consider a single electron in the lowest SL miniband EAl(b). In the
presence of an electric field parallel to the SL axis, the electron is accelerated
432 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Figure 3.76: Exper-


imental current-voltage 300K
characteristics of various
GaAs/AlAs SLs with the
electric field parallel to >-
the SL axes. (After 5 -
Sibille, Palmier, Wang,
and Mollot, 1990.)

2 4 6 0 3
BIAS (VI

until it reaches the inflection point of Eil(k). Thereafter, its effective mass
becomes negative, causing deceleration instead of acceleration. Since the
electron is inevitably scattered by phonons and impurities, its velocity takes
a stationary value in a dc field. The latter increases with increasing field
strength E if the corresponding k-value is below the inflection point, and
it decreases with increasing E if the k-value of the electron is above this
point. However, a superlattice contains a gas of many electrons, populating
the miniband E:(k). For low field strengths, most of the electrons have low
k-momenta and positive effective masses. Thus, the total current density of
the gas increases if E is further increased. Simultaneously, the electron gas
is heated up by the electric field, which implies that more electrons populate
k-points with a negative effective mass and less with a positive one. This
reduces the current increase if E raises further because a larger subgroup of
electrons is decelerated instead of being accelerated by the field increase. If
the subgroup of electrons having negative effective mass is sufficiently large,
further increase of the field strength will actually decrease the current. This
explains the negative differential drift velocity seen in Figure 3.75.
An example of an optical experiment is shown in Figure 3.77. There,
the absorption spectra of GaAs/Al,Gal-,As QW structures of different wen
widths are shown. The staircase-like shape of these spectra reflects the DOSs
of the electron and hole subbands of a QW which above have been shown
to be staircase-like as well. Also other features of the QW band structure
observed above, as the widening of the fundamental energy gap and the
shift of the steps towards higher energies with decreasing well width, can be
clearly seen in the spectra of Figure 3.77. The enhancement of absorption
at the step edges, which correspond to optical transitions from the tops of
hole subbands to the bottoms of electron subbands, is due to the Coulomb
3.8. Macroscopic electric fields 433

Photon energy ( e v ) ---w

Figure 3.77: Absorption spectra of GaAs/Al,Gal-,As QWs of various well widths.


(After Dingle, Wiegmann, and Henry, 1974.)

interaction between electrons and holes (excitonic effects).


Beside absorption spectra, photoluminescene spectra are extensively used
for the experimental characterization of semiconductor microstructures. In
photoluminescene, one observes mainly radiative transitions between the
electron and hole ground state levels of an (undoped) QW because only
these are considerably populated. Due to the enlarged energy separation of
these levels, a QW emits light at shorter wavelength than the corresponding
bulk material does. In the case of GaAs this effect is particularly striking
- the bulk crystal emits radiation in the non-visible infrared region, while a

GaAs-quantum well of 50 A width emits visible red light.

3.8 Macroscopic electric fields


Macroscopic electric fields in semiconductors are the central focus of many
device applications and give rise to important physical phenomena in these
materials. For instance, if one applies an external electric field to a semicon-
ductor sample, the response yields information about the band structure,
impurity states and other microscopic properties of the material. The elec-
tric transport properties and electro-optic effects are particularly well suited
434 Chapter 3. Electronic structure of semiconductor crystals with perturbations

to extract this information. Often, electric fields in semiconductors need


not, however, be applied externally as they are already present internally,
either because of spatially inhomogeneous doping such as in the case of a
pn-junction, or because of other spatial inhomogeneities as surfaces or in-
terfaces. In the following theoretical description, the source of the electric
field will play a role only inasmuch as we assume that the field is spatially
uniform. In practical terms, this means that the field should not change
appreciably within a periodicity region, i.e. within a characteristic length of
G x a. Since the limit of the infinitely extended semiconductor is effectively
reached in good approximation with G M 100 the characteristic length need
not to be larger than about 0.01 ,urn. Fields which change only little over
this very small distance can be considered as homogeneous. In addition to
excluding spatial inhomogeneities of the electric field, we will also exclude
temporal changes in our discussion below. This only means that the fre-
quencies of these changes should be small compared with the characteristic
frequencies of the electrons of the semiconductor.

3.8.1 Effective mass equation and stationary electron states


A semiconductor in a homogeneous external electric field E represents a
perturbed crystal in the sense of sections 3.1 and 3.2. The presence of such
a field can be described by adding a perturbation potential V(x) to the
one-electron Hamiltonian H of the ideal crystal. This potential is defined
as the difference of the energy of a crystal electron in the presence of the
electric field and without it, and thus is given by

V(x)= e E . x. (3.294)

As in the case of point perturbations considered in sections 3.4 and 3.5,


the perturbation potential V(x)of equation (3.294) does not possess lattice
translation symmetry. Moreover, V(x) diverges at infinity. The latter fact
implies that, unlike to the case of point perturbations, an infinite crystal
in a homogeneous electric field cannot be replaced by a perturbed supercell
whose periodic repetition forms an unperturbed supercrystal. The crystal
in an electric field has to be treated as what it in fact is, namely an infinite
system with 2-dimensional lattice symmetry perpendicular to the field, and
no lattice symmetry along the field direction.
If the field strength E is not too large then the perturbation potential
(3.295) fulfills the smoothness condition of effective mass theory of section
3.3, which here reads
e I E I a << E g . (3.295)
With this condition fulfilled, the effective mass equation (3.53) is applicable.
Here it has the form
3.8. Macroscopic electric fields 435

{ E y ( - i V ) + e E . x } F V ~ , ( x )= EvFv~,(x). (3.296)
We consider this equation in the particular case of an isotropic and parabolic
band. Assuming the z-axis of our Cartesian coordinate system parallel to
the field direction, the electron motion in 2- and y-directions is that of a free
particle. Thus the envelope function F,(x) can be written as

F~E,(x)E F ~ E , ~ , ~ (=
X)e i k ~ ~x,,
. x (2).
~ (3.297)
where x l and k,l denote, respectively, the components of x and of the
wavevector perpendicular to the field. The z-dependent envelope function
factor X , , ( t ) obeys the equation

X,,(z) =E ~ X , ~ ~ ( Z ) , (3.298)

with
li2
E , = E + -2m
k:l. (3.299)

For the conduction and valence bands, in particular, equation (3.298) reads

(3.300)

(3.301)

The second equation can be interpreted as the effective mass equation for
a hole with (positive) mass Im:l and energy (-.,). As the equation shows,
the charge of the hole is positive, i.e. fe, in contrast to the electron with
charge -e.
The spectra of energy eigenvalues of equations (3.300) and (3.301) are
continuous, and vary between -co and fco as a direct consequence of the
unrestricted motion of the electron or hole in the field direction. This allows
one to use the energies E , as continuously varying quantum numbers of the
stationary states X V C v ( z )and
, to assume normalization in terms of Diracs
&function with respect to these energies, thus

The solutions of the corresponding energy eigenvalue problems for electrons


and holes read, respectively,
436 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.303)

(3.304)

with

(3.305)

as so-called electro-optic frequency, and with Ai as the Airy-function of first


kind. Note that stationary states X,,(z) of conduction and valence band
electrons exist at all energies ey between --oo and +m in the presence of an
electric field rather than only at energies in the corresponding e n e r a bands
without such a field.
Assuming E = lo4 Vlcm and m: = 0.1 m , one obtains for 0 , a value
of about 1013 s - l , and for he, a value of 10 m e V . The profiles of the
two envelope functions X E C ( z )and X , , ( z ) are shown in Figure 3.78 for
eC = eV = 0 and Bc = 0,. The electron function decays exponentially with
increasing z for e E z > -Eg,and oscillates with decreasing z for e E z <
-Eg. On the decaying side, the total energy E g of an electron is smaller
than its potential energy e E z . Thus, this region is classically forbidden, it
represents a potential barrier for electrons. The envelope function is non-
zero only because electrons tunnel into this barrier. The envelope function
for holes behaves similarly, it decays within the hole barrier with e E z < 0,
and oscillates in the classically allowed hole region with e E z > 0.
The Schrodinger equations (3.300), (3.301) and their solutions (3.303)
and (3.304) describe, respectively, stationary electron and hole states in an
electric field. Electrons or holes in such states do not carry an electric
current, more strictly speaking, the expectation value of the current density
operator in the field direction vanishes. Although this does not mean that
a current flow cannot be described at all by means of stationary states (if
the electron system is characterized by a statistical ensemble with respect
to stationary states, then non-zero non-diagonal elements will occur in the
expectation value of the current density operator), one may suspect, however,
that in addition to stationary states there will be yet other states which are
better suited to the condition of a non-vanishing current. Such states do
in fact exist, they are not eigenstates of the Hamiltonian but rather are
non-stationary states cp,(x, t ) which devebp in time not only by means of a
timedependent phase factor. They correspond to the classical trajectories
of an electron in an electric field, with momentum growing linearly in time.
3.8. Macroscopic electric fields 437

0.4

C
0.3
.-
u
0
(D
g3 0.2 rn
Lc

g 0.1
-e
N

-0 P
g
W
0.0
c
2
rt
v)
Y

-0.1

-0.2

-Eg/eE 0 Z

Figure 3.78: Envelope functions of stationary electron and hole states in an electric
field, taken at energy cC = E, = 0. Oc = OU = 8.

3.8.2 Non-st at ionary states. Bloch oscillations


The envelope functions F,(x,t)of these states are time-dependent, so one
has

$v(x, t ) = ~ V O ( X ) F V ( Xt ,) . (3.306)

To determine F,(x, t ) ,a time-dependent effective mass equation is required.


This may be obtained from the corresponding time-independent equation
(3.296), wherein the energy E, on the right hand side is replaced by the
operator i b ( a / a t ) . The resulting equation reads

a
+
{&,(--iV) eE . z} Fv(x,t ) = iTi-F,(x, t ) .
at
(3.307)

In solving this equation we assume that the electron was in a Bloch state
(p,k(x) at t = 0. Then the initial condition for F,(x,t ) may be stated as

FJX, t = 0) s F,/C(X, t = 0) = -.ik.x (3.308)


G '
Here, the envelope function at t = 0 has been chosen such that its normal-
ization integral with respect to the infinite interval over which the crystal
in field direction extends, is given by a &function with respect to k. The
solution F,k(x, t ) of equations (3.307) and (3.308) is given by
438 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.309)

with
kt = k - ZEt. (3.3 10)
fi
This is readily verified by a straightforward calculation. Owing to the en-
velope function (3.309) and the previously made approximation U& = Uuk,
the total non-stationary wavefunction & , k ( x , t ) evolving from a Bloch state
(pvk(x)at t = 0 remains a Bloch state for t > 0 also, however, with a
time-dependent quasi-wavevector kt,

(3.311)
This important result may also be established without approximating the
Bloch factor U v k by u,o as is done in effective mass theory. The only ap-
proximation needed is the neglect of interband transitions induced by the
electric field.
Equation (3.311) states that the quasi-wavevector k of a Bloch state in
the presence of an electric field changes with time in just the same way as
does the ordinary wavevector of a free electron. Differing from the latter,
the quasi-wavevector is restricted to the first B Z . However, relation (3.310)
results in kt-values which can also lie outside the first B Z . This means
that in the above derivation we inadvertently changed from the reduced
to the extended zone scheme. According to the general considerations of
Chapter 2, one may recover the description in terms of the first B Z by
subtracting a suitable reciprocal lattice vector K(t). For electric fields in
symmetry directions of the first B Z , there will be a particular point in time
t = T(E) at which, with increasing t , the reduced wavevector kT - K(T)
equals the wavevector kt at t = 0 for the first time. This means that the
time-dependent Bloch state (3.311) returns, at t = T , to the Bloch state
at t = 0 (not counting a phase factor). The same holds for time points
2T,3T etc. The time development of Bloch states in an electric field is
therefore periodic, hence the term Bloch oscillations. The period T of these
oscillations may be obtained from (3.310) as

(3.312)

The Bloch oscillations affect not only the wavefunctions but also the perti-
nent energies E,(kt), meaning that Bloch electrons execute periodic motion
within their energy bands.
Oscillations also occur in the velocity of Bloch electrons, defined by the
quantum mechanical expectation value < ( d x / d t ) >= (qJvhI(dX/dt)IVvk,)of
3.8. Macroscopic electric fields 439

the velocity operator ( d x l d t ) with respect to the timedependent Bloch func-


tion Pvk,. using the identity (dxldt) = ( l / m ) p , the velocity < ( d x / d t ) >
follows by means of the previously derived relation (2.193) for the expecta-
tion value ((p,hJpJ(p,h)of the momentum operator p. The timeaverage of
the velocity < ( d x / d t ) > obtained in this way, taken over a Bloch period,
turns out to be zero. The same result holds for the timeaveraged current of
the electrons of a crystal in an external electric field, it also vanishes. That
this is actually not the case is due to perturbations of the Bloch oscillations
by collisions of electrons with phonons, impurities and other point perturba-
tions. To prove this statement we consider a field strength of lo4 V / c m and
a primitive reciprocal lattice vector K typical of semiconductors. The Bloch
period T is of the order of magnitude lo-'' s in such circumstances. The
time rv between two collisions is substantially smaller than T , typically of
the order of magnitude s. Thus, an electron starting a Bloch oscilla-
tion will soon be scattered and its momentum will be distributed randomly
over all k-space. In other terms, no periodic motion can develop in the pres-
ence of collisions. We conclude that collisions are not only responsible for
the fact that the current of a free electrons, which would otherwise be in-
finitely large, remains finite, but also for the fact that an otherwise vanishing
average current of a Bloch electron does not actually go to zero.
There is still another reason that Bloch oscillations cannot be observed
in ordinary semiconductor crystals. It is due to the fact that the time depen-
dent Bloch states (3.311) decay in time by themselves since the exponential
factor oscillates with increasing frequency. This may easily be seen for quasi-
wavevectors k close to a critical point. For zero wavevector component k,
in the field direction, the phase of the exponential contains a term (1/3)e:t3
(0, is the electro-optic frequency of equation 3.305). If (1/3)0:t3 is consider-
ably larger than 1, the exponential oscillates so fast that the time dependent
Bloch states (3.311) average out to zero almost completely. One may say
that these states have a finite lifetime 0;' due to the electric field. For not-
too-large field strengths, e;' is small compared with the period T of Bloch
oscillations, just as the time T, between two collisions above was found to be
small with respect to T . This implies that, in actual crystals and under nor-
mal conditions, Bloch oscillations are not observable even without collisions,
just due to the finite field induced lifetime.
For the artificial superlattices considered in the previous section 3.7, the
primitive lattice vectors and thus the periods T are much shorter than those
in natural crystals. In such circumstances T may become larger than T~ and
OF1,and Bloch oscillations may in fact be observed. The negative differential
electric conductivity of a superlattice parallel to its axis, reported in section
3.7, also may be understood as a manifestation of Bloch oscillations.
If one considers only k-vectors close to critical points of the band struc-
ture, then the energy bands E,(k) can be taken in parabolic approximation.
440 Chapter 3. Electronic structure of semiconductor crystals with perturbations

This yields, approximately,

(3.313)

With the neglect of collisions, an equation similar to Newtons law of motion

m t -d 2 < x > = -eE, (3.314)


dt2
follows from (3.313). It differs from the ordinary Newtons equation in that
the free electron mass is replaced by the effective mass of the respective band
v. For electrons from the valence band the effective mass is negative. The
pertinent particles with positive mass, the holes, behave like particles with
positive charge e.

3.8.3 Interband tunneling


In the effective mass equation (3.296), the possibility of electric field in-
duced coupling between the valence and conduction bands is neglected. In
reality such coupling exists due to the non-vanishing interband matrix ele-
ments of the potential V(x) = e E . x. Taking account of these off-diagonal
elements yields a description of electrons undergoing quantum mechanical
transitions between the valence and conduction bands. For reasons of energy
and momentum conservation, the initial and final state energies E,, Ec and
wavevector components k,l, k,l perpendicular to the field have to be equal
in such a transition. According to equation (3.299) this means E, = cC. The
quantum mechanical probability W , for field induced interband transitions
is proportional to the overlap integral between the two envelope functions
F,,vkvl(x) and FECkci(x) with identical values for E, tC and k w l ,kc-. If,
in particular, transitions of electrons between the two band edges, i.e. with
E, = ec = 0 and k ,l = k,l = 0, are considered, the overlap integral has
to be taken with respect to the product Ai([E,+ e E ~ ] / A e , ) A i ( [ - e E z ] / A e , ) .
This integral is non-zero since valence and conduction band states having
the same energies E, = eC = 0 differ from zero simultaneously for almost all
values of z , as may be seen from Figure 3.106. The reason for the non-zero
overlap is the tunneling of the valence and conduction band electrons into
their respective barriers. This explains why field induced interband transi-
tions are referred to as i n t e r b a n d t u n n e l i n g (the term Z e n e r t u n n e l i n g is also
used).
+
Performing the overlap integral of A i ( [ E , e E z ] / ~ e , ) A i ( [ - e E z ] / A e , )
with respect to z one arrives at an expression proportional to the square of
the Airy-function Ai3(E,/he,) where 8, is a frequency analogous to the
electro-optic frequencies e , , of equation (3.305), formed here, however with
+
the reduced effective mass mT, = mZ,m:/(mz m:) of electrons and holes
3.8. Macroscopic electric fields 441

instead of their simple effective masses m,*,,. One thus obtains the interband
transition probability W , as

W,, oc A i 2 (z) .
For realistic field strengths E, is measurably larger than TLB,,.
(3.315)

Therefore,
the asymptotic approximation

(3.316)

for the Airy-function applies. To understand what this result means phys-
ically we may argue as follows. Expression (3.315) for the transition prob-
ability W , was derived using the stationary electron states in an electric
field of subsection 3.8.2. The same expression follows if one uses the non-
stationary states of subsection 3.8.3. In this, the electro-optic energy Ti@,,
may be understood as the energy uncertainty of an electron-hole pair due
to its finite lifetime in the presence of an electric field. Without such an
uncertainty interband tunneling would not be allowed by reason of energy
conservation: the final state in the conduction band would differ in energy
from the initial state in the valence band by the gap energy E,. The energy
uncertainty relieves the need to satisfy energy conservation and permits the
interband transition probability W , to be nonzero. The non-stationary
approach also provides a rough estimate of the proportionality factor in re-
lation (3.315). The probability W,, for an electron to tunnel per second
from the valence into the conduction band at k,l = kl = 0, may be ob-
tained from Ai2(E,/7iB,,) by multiplying this expression with the frequency
1/T of Bloch oscillations. This is plausible since this frequency indicates
how often a valence band electron, per second, reaches the upper band edge,
from which it can tunnel into the conduction band. It is just the probabil-
ity for this tunneling step which the Airy-function expression A i 2 ( E g / W , )
represents. The complete expression for W , thus reads

1
W , = -Ai2
T
(2). (3.317)

Although the interband matrix element (p,kJeE.xlp,k) of the perturbation


potential V'(x) does not enter this expression explicitly, as one would ex-
pect, it occurs implicitly through the lattice constant a in T , which has the
same order of magnitude as (p,k[xIp,k). According to expression (3.315),
the tunneling probability W,, grows larger as the energy uncertainty hOCw
approaches the gap energy E,. It is small if the condition

fit', << Eg (3.318)


442 Chapter 3. Electronic structure of semiconductor crystals with perturbations

holds. In typical cases this condition is fulfilled sufficiently well for field
strengths of the order of magnitude lo5 V / c m , while it is definitely not valid
for field strengths close to the inner-atomic fields of 10' V / c m . At least
for such fields interband transitions are important. By means of tunneling,
holes are generated in the valence band and electrons in the conduction band.
These diminish the resistivity of a semiconductor material and initialize other
processes such as carrier heating and impact ionization of the valence band
which may finally result in electric breakdown.
The inequality (3.318) transforms into the condition (3.295) for the ap-
plicability of effective mass theory if the gap energy Eg is formally identified
with (h2/2m5,)1/a2. The latter energy represents an average energy gap in
effective mass approximation. It has the same order of magnitude as E,, so
that the two conditions (3.318) and (3.295) are equivalent.

3.8.4 Photon assisted interband tunneling


As pointed out in Chapter 1, transitions of electrons from the valence into
the conduction band can also be caused by photons provided their energy
JLw exceeds the gap energy E, of the semiconductor. In the presence of
an electric field this condition no longer needs to be fulfilled, owing to the
tunneling of electrons and holes into their respective barriers, effectively low-
ering the gap (see Figure 3.79). The optical interband transition probability
W z t between the two band edges is proportional to the overlap integral of
the envelope functions FVevkvl(x) and FCL,kCl (x)with identical values for
k,l, kl,just as in the case of field induced interband transitions, but, un-
like the latter case, with energies eC = e,+hw. This means that W s t follows
from W , in equation (3.315) by replacing E, with E, - hw. One therefore
has

(3.319)

In an external electric field, the optical transition probability and, hence, the
optical absorption coefficient, is also non-zero for photon energies below the
gap. It decays exponentially as hw decreases. This phenomenon is called
the Franz-Keldysh effect (Franz, 1958; Keldysh, 1958; Boer, Hansch and
Kummel, 1959). Above the gap, the absorption coefficient with electric field
oscillates with photon energy about the absorption coefficient without field,
and we have Franz-Keldysh oscillations (Tharmalingan, 1963).
Field induced changes occw not only in the absorption coefficient, but
also in other optical constants including the real part of the complex refrac-
tive index. Besides the changes at the fundamental absorption edge consid-
ered above, one also has changes of optical constants at any higher van Hove
singularity of the joint density of states of the interband energy Ec(k)-E,(k)
3.9. Macrmcopic magnetic fields 443

- ( E,-ho)/eE 0 2

Figure 3.79: The same envelope functions as in Figure 3.78, but with the energy
of the electron exceeding that of the hole by the photon energy iiw.

(Aspnes, 1967; Enderlein and Keiper, 1967). This fact is exploited in elec-
trorefiectance spectroscopy, in which field induced changes of the reflectivity
of a semiconductor sample are measured by means of a modulation tech-
nique. Photon energies which give rise to strong electroreflectance signals
correspond to optical transitions at critical points of the interband energy
(see Figure 3.80). By measuring the electroreflectance spectra of semicon-
ductors one obtains experimental data relating to their valenceconduction
band separation at critical points. These can be used as input parameters
for empirical band structure calculations (Cardona, 1969) as well as for the
characterization of semiconductor heterostructures, including superlattices
and quantum wells (Pollak, 1994; Enderlein, 1996).

3.9 Macroscopic magnetic fields


Like electric fields, macroscopic magnetic fields also give rise to effects in
semiconductors which provide experimental data concerning effective masses
and other microscopic properties of electrons and holes. Transport proper-
ties, particularly galvanomagnetic phenomena such as the Hall effect, mag-
netoresitivity and cyclotron resonance, as well as magneto-optic properties,
are of particular importance in this context. The basis for theoretical un-
derstanding of these phenomena are the stationary one-electron states of a
semiconductor crystal in the presence of a magnetic field. For most purposes
444 Chapter 3. Electronic structure of semiconductor crystals with perturbations


-
-121 I I I I I I I
1.0 1,5 20 25 30 35 400 4s
Energy ( e V )

Figure 3.80: Electroreflectance spectrum of Germanium. (After Seraphin and


Hess, 1965.)

this field can be assumed to be homogeneous in space and time. Just as in


the case of external electric fields, this again only means that the magnetic
field is approximately constant on length and time scales which are large
compared with corresponding microscopic length and time scales. In the
case of cyclotron resonance, the microscopic time scale is given by the ro-
tation period of an electron about the magnetic field axis which typically
lies in the G H z range. Thus, temporal changes of the magnetic field in the
M H z range are still admissible. Below, we consider spatially and temporally
constant magnetic fields H exclusively. The vector of the pertinent magnetic
induction vector will be denoted by B, as usual.
Since the interaction of an electron with a magnetic field cannot be char-
acterized by a scalar perturbation potential, but must be described by the
vector potential A of the magnetic induction B, the effective mass theory
developed in section 3.3 is not directly applicable to this case. Whether an
effective mass equation can be derived at all with a magnetic field, and what
it will look like if it is feasible, must be explored separately. This matter will
be addressed in the following subsection.

3.9.1 Effective mass equation in a magnetic field


Initially, spin will be omitted from our considerations (it will be taken into
account later). In these circumstances, the role of the magnetic field in the
Hamiltonian is fully subsumed in the kinetic energy operator (p2/2m) with
the canonical momentum operator p replaced by the kinetic momentum
+
operator p (e/c)A(x). The vector potential A(x) is determined by the
magnetic induction vector B, save for the gradient of an arbitrary gauge
3.9. Macroscopic magnetic fields 445

function which we choose so that A ( x ) has the form

A(x) = A[B x XI. (3.320)


2
This choice of gauge guarantees that p commutes with A(x), so that the
order in which p and A(x) are multiplied in the expression [p (e/c)A(x)I2 +
is inconsequential. The strength of the magnetic field will, as in section
3.8 on electric fields, be limited by the condition that the associated vector
potential changes only little over a unit cell in comparison with the energy
gap. In this sense, the vector potential is presumed to be smooth.
The Schrodinger equation of a spinless crystal electron in a magnetic field
reads

(3.321)

As in the derivation of the effective mass equation for perturbing scalar po-
tentials in section 3.3, we use the k . p-perturbation theory, i.e. we represent
the Schrodinger equation in the approximate Bloch basis Iuk)', whence

u'k'
'(vk(H + LmcA ( x . p ) + l
(:) 2
A2(x)IvAk')l '(vhk'lg) =E '(vkI$).
(3.322)
with H = (p2/2m) + V ( x ) as Hamiltonian of the crystal without a magnetic
field. Initially, we consider a non-degenerate band extremum at k = 0, and
assume the form (3.60) for '(vkl$). In order to satisfy the Schrodinger
equation (3.322) with this Ansatz, the component with u = vo must obey
the relation

k'
l(vok(H + emAc ( x ) . p + 1
f) A2(x)(uok')'Fw(k') = EF,,(k).
(3.323)
Components '(vkl$) with v # vo must vanish according to equation (3.322).
To verify this we consider the matrix elements '(vklHlvok')' of the unper-
turbed Hamiltonian H , which were already evaluated above. The expression
(3.63) derived there is applicable here as well. Accordingly, t-hematrix ele-
ments of H with v # vo vanish.
Because of the assumed smoothness of the vector potential A(x), its
matrix representation '(vkJAlv'k')' is approximately diagonal with respect
to band indices, whence
446 Chapter 3. Electronic structure of semiconductor crystals with perturbations

The diagonality of the matrix for A extends to the matrix for A2, as well,
and therefore one also has

(vkl A2Ivk) M 6
1
,( klA21k) . (3.325)
What remains to be evaluated in the Schrodinger equation (3.323) now is
only the matrix of the operator ( e / m c ) A .p. This operator has the same
structure as the k . p-perturbation operator. Just as the matrix represen-
tation of the latter with respect to the approximate Bloch basis (vk) is
band-diagonal in second order perturbation theory, the same also holds for
the matrix representation of the operator ( e / m c ) A . p. We have

(3.326)

This relation finally confirms the vanishing of the v # vo-components of the


left hand side of Schrodingers equation (3.323) and, thus, of (vklq) with
v # vo.
The band-diagonal elements (vokl e / m c A .plvok) will now be rewritten
in a more convenient form. We proceed in two steps. In the first step, we
determine the matrix elements of the operator ( e / m c ) A . p between the
Luttinger-Kohn functions (vokO) Ivok) which enter the first-order Bloch
functions Ivok) of equation (2.334) as unperturbed parts. In the second
step we consider the first-order corrections.
First step
Since
1
A.p= -L.B (3.327)
2
with L = [xx p] as angular momentum operator, we have

(vokOleA. pIvok0) = pB(vokO(Ti-LIvokO) . B. (3.328)


mc
Here p~ = (eTi/2mc) denotes the Bohr magneton. The matrix elements
of L can be rewritten by means of Heisenbergs equation of motion which
approximately yields

(3.329)

Here all terms depending on the magnetic field have been neglected since
they only give rise to quadratic terms with respect to B in the matrix ele-
ments (3.328). Substituting relation (3.329) in expression (3.328) for these
elements, and using the notation L , = zap^ - @pa for angular momentum
components (c.Pr signifies the ordered triplet z y z or a cyclical permutation
of it), we obtain
3.9. Macroscopic magnetic fields 447

(3.330)
In this, we used the fact that the interband matrix elements of the momen-
tum operator are diagonal with respect to k and k. The matrix elements
(vokO)L,lvokO) do not depend on k, so we may use the relation

The quantity L, is an antisymmetric tensor or, in other terms, a pseudovec-


tor L,. This antisymmetric tensor bears a close relation to the reciprocal
effective mass tensor M&La of equation (2.337). Apart from a constant fac-
tor, the two quantities may be understood as real and imaginary parts of
the same complex tensor, namely

(3.332)

In section 3.3 we found that the symmetric tensor M,Lp describes the k-
dependence of the band-index- and k-diagonal Bloch matrix elements of the
Hamiltonian (see equation (3.63)). Here we find, that the antisymmetric
tensor L, determines the B-dependence of these elements. The latter
evidently describe the interaction of the magnetic moment of the orbital
motion of a Bloch electron with the magnetic induction vector B. In order
that it differ from zero, the angular momentum of the Bloch state IvoO)
must differ from zero, too. In principle, this is possible because the motion
of a Bloch electron cannot in general be reduced to a translation, since it
contains a rotational part about the atoms which form the crystal. The
tensor L, measures the average angular momentum of this rotation in the
Bloch state 1~00). In cubic materials it vanishes for symmetry reasons, at
least as long as only non-degenerate Bloch states IvoO) at the center of the
first BZ are considered and spin and spin-orbit interaction are omitted from
consideration, as we do here. The situation is comparable with that for
s-states of free atoms.
Second step
Beside the k-diagonal term discussed above, which describes the coupling of
the magnetic field to the rotational part of the motion of a Bloch electron, the
448 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Hamiltonian matrix of (3.323) still contains terms which are non-diagonal


with respect to k, k and give rise to a coupling of B to the translational part
of this motion. They occur through the formation of the matrix elements
l(vok)(e/rnc)A.p)lvok), where the perturbed part [Ivok) - IvokO)] of the
approximate Bloch functions Ivok) (which was previously omitted) is taken
into account. The complete result is given by

6kk,h-1pBB. L, +-1
2m
[k,e(k(Aa(k) +
kp(kja,jk)] . (3.334)
4
Summing the three terms (3.63), (3.334) and (3.335) obtained in rewriting
the Schrodinger equation (3.323), and simultaneously transforming from k-
space to coordinate space, the following equation results for the envelope
function Fvo(x):

For a non-degenerate isotropic parabolic band the magnetic moment L,


vanishes in the spinless case considered here, as already mentioned. Then
equation (3.335) takes the form

(3.336)

This equation changes if spin and spin-orbit interaction are taken into ac-
count. The assumption of a non-degenerate band vo will be adopted, initially.

Spin a n d spin-orbit interaction

Considering the role of spin and spin-orbit interaction, the Schrodinger equa-
tion (3.321) may be written as

where g is the gyromagnetic ratio of a free electron (g = 2.0023). Using the


same procedure employed in the spinless case, i.e. by first representing equa-
tion (3.337) in the spin-dependent Luttinger-Kohn basis, then transforming
3.9. Macroscopic magnetic fields 449

it into the spin-dependent approximate Bloch basis Ivak)' = Ivk)'lo), and


finally transforming it back to coordinate space, the effective mass equation
for an isotropic parabolic band follows as

= EF,(x,s). (3.338)
In obtaining (3.338), we used the fact that the diagonal elements '(voaklH,,
Ivoa'k)' of the spin-orbit interaction operator vanish at k = 0 for symmetry
reasons, while the non-zero k-linear terms in these elements are negligible.
The matrix elements L, of the angular momentum operator emerge from
the corresponding expressions (3.330), (3.331) without spin-orbit interaction
by replacing p in the latter with the spin-dependent operator ?i of equation
(2.353), obtaining

In accordance with their definition, the L,, are matrices in the two-compo-
nent spin space. In regard to their transformation properties in coordinate
space, they are components of a pseudovector. Apart from a constant factor,
there is only one such pseudovector in the case of systems with the symmetry
group oh, namely the vector of Pauli's spin matrices a. This means that
L, is a multiple g,Acr' of cr'. The constant gvo is determined by the matrix
elements of (3.339). With + 1
= guo 29, we may write
1
h-lLm + 2g cr' = i g & d . (3.340)

Using this expression, the effective mass equation (3.338) takes the form

The factor gEo is called an effective g-factor. Just like the effective mass,
it represents a characteristic of an energy band vo at a particular critical
point. In Table 3.11 experimental g:,-values for the conduction band min-
imum of several diamond and zincblende type semiconductors are listed.
As one may see, all values gE are less than 2, the g-factor of the free elec-
tron. The deviation from 2 is determined by the induced magnetic moment
of the orbital-motion component of a Bloch electron. As in a free atom,
this moment is always negative, in other words it results in a diamagnetic
450 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Table 3.11: Effective g-factors gE of electrons and magnetic interaction constants


and q of holes in diamond and zincblende type semiconductors. (After Landoldt-
IC

Bornstein, 1982.)

GaAs GaSb InSb CdTe ZnSe

-0.42 3.41 1.20


0.06 0.04 0.13 0.15

contribution to the total magnetic moment of the Bloch electron. For Si,
the effective electron g-value is close to 2, indicating that the diamagnetic
contribution is small in this case. A negative effective g-value and hence a
larger diamagnetic contribution occurs in Ge. This is to be expected since
the diamagnetic moment is determined by the average angular momentum
L
, which according to expression (3.339), and similar to the effective elec-
tron mass of expression (2.338), increases if the gap decreases. The same
behavior of the effective g-factor for decreasing gap is observed among the
111-V compound semiconductors: InSb, the material with the smallest gap,
has the largest negative value of g:.
To describe the top of the valence band of diamond and zincblende type
materials, we need an effective mass equation for degenerate bands, and will
address this below.

Degenerate bands

We consider the valence band of a diamond and zincblende type semicon-


ductor in the vicinity of the BZ center at k = 0 . Spin and the spin-orbit
interaction will be omitted initially. Then the degeneracy is %fold. The
Bloch states at k = 0 are IvmO),m = 2,y, z , and the pertinent Luttinger-
Kohn functions read (vmkO). The eigenfunction $(x) is given in terms of
*
the envelope function vector F , (x)= (Fwz(x), F w ( x ) , F,,(x)) by equation
(3.67). The elements MGip of the effective mass tensor are 3 x 3 matrices
04
D with elements DZ:,. The same holds for the components L, of the
effective angular momentum of equation (3.333) which may be expressed as
0
matrices L~ with elements
3.9. Macroscopic magnetic fields 45 1

It follows that the envelope function vector of the degenerate valence band
in a magnetic field satisfies the equation

where the abbreviation

(3.344)
has been used for the angular momentum term. Like the first tensor in equa-
9
tion (3.343), the second tensor H M can also be determined, apart from cer-
tain constants, by means of the point symmetry of the crystal. The method
of invariants, which was formulated in section 2.7 for the k-dependent term
9
of the Hamiltonian, is also well suited for the B-dependent operator H M .
The only change is the replacement of the wavevector k by the magnetic
induction B. To apply this method to the I?;, valence band of a diamond
type semiconductor, one needs a complete set of 9 linearly independent basis
9 9 9
vectors in the 3 x 3-matrix space formed from the three matrices I z rI y rI z
of angular momentum quantum number j = 1 and from the six products
of these matrices which transform according to the irreducible parts of the
product representation x I?;, = I?l +I?12+I':,+I';, (for the explicit form
of the angular momentum matrices see equations (A.143) of Appendix A).
Since the magnetic field B, as a pseudovector, belongs to the representation
I?:, of the point group oh, only subsets of matrices transforming according
to the irreducible representation ,:?I of Oh are allowed to form an hvari-
9
ant H M under the point group operations of oh. The subset of matrices
which transforms according to the representation I?:, is the pseudovector
9 9 9 9 9

-
I = (Iz,I y rIzLitself. Thus the scalar product (BI I ) emerges as the only
invariant, and H M follows as

(3.345)
with K being a constant which has to be determined from band structure
calculation or by means of experimental investigation. The same result for
9
H M follows for the I'l5-valence band of zincblende type semiconductors.
The above consideration shows that, in the case of degenerate bands,
there is coupling of the orbital motion to the magnetic field which does not
occur for non-degenerate bands. The source of it is the non-vanishing of
angular momentum matrix elements between the degenerate valence band
states. Of course, the average angular momentum of all valence band states
also vanishes in the degenerate case. However, the various angular mo-
mentum states, averaging to zero in the absence of a magnetic field, are
452 Chapter 3. Electronic structure of semiconductor crystals with perturbations

differently affected by the magnetic field so that a magnetic splitting of the


r;, or r15 valence band occurs, comparable to that of a p-level in the case
of a free atom.
If one adds the effects of spin and spin-orbit interaction, then the I'b,
band of diamond type semiconductors develops into the r,f and I:' bands.
The effective mass equation for the I?$ band without a magnetic field is given
by (3.74). Two changes are necessary in the presence of a magnetic field:
First, in the operators Q, R , $, f of equation (3.74), the replacements 8, --.$
d , + ( i e / h c ) A , , Q = 2,y, z , have to be made. Second, an angular momentum
($
term H M has to be added. The latter can again be obtained by means of the
method of invariants. To apply this method in the present case, one must
find a complete set of 16 basis vectors of the 4 x 4-matrix space formed from
the three angular momentum matrices I ~ly,
( $ @ @
4
, I z of quantum number of
equation (A.144) and their products which transform in accordance with the
irreducible parts of the product representation r$ x I?$ = rl+ rz I'12 + +
21'i5+I'k5 of oh. As in the case of the I';, band without spin, invariants linear
in B are possible only with subsets matrices of I'i5-symmetry. Differing from
9 u u *
the spinless case, there are now two such subsets, namely I = ( Ix,I y rI z ) ,
030303 *3
and also the triplet (Ix,I,, I z ) . We denote it symbolically by I . The
*
additional term H M in the Luttinger-Kohn Hamiltonian which describes
the coupling of the magnetic field to the magnetic moment of the I'$-valence
band states thus reads

(3.346)

where K and q are again constants which must be determined either from
band structure calculations or experimentally. That the presence of spin
involves not only K , but also a second constant q, is understandable because
the orbital and spin motions have different gyromagnetic ratios. In most
(z
cases q is negligibly small (see Table 3.11). The result (3.346) for H M which
has been derived here for diamond type semiconductors, holds unchanged
for materials with zincblende structure also.

3.9.2 Solution of the effective mass equation


We demonstrate the solution of the effective mass equations in a magnetic
field in the simplest possible case, that of a non-degenerate band without
spin. In this, we take the magnetic field in the z-direction and employ the
vector potential in the form A = (--By, 0,O). This gauge differs from that
in expression (3.320) used in deriving the effective mass equation (3.336).
Unlike (3.320), the new gauge lacks axial symmetry with respect to the
magnetic field. The effective mass equation (3.336) also holds, however,
3.9. Macroscopic magnetic fields 453

because for this gauge the vector potential A = (--By, 0,O) also commutes
with the momentum operator p. The same holds for the more complex
effective mass equation (3.343) for degenerate bands without as well as with
spin. Using the present gauge, the effective mass equation (3.336) for a
non-degenerate band without spin takes the form

2-
am:, [
(Pz + $Y)2 +Pp +pq F d X ) = ( E , - E,o)F,(x). (3.347)

Since the Hamiltonian in (3.347) commutes with the translation operators


for arbitrary displacements in the x- and z-directions, its eigenfunctions can
be chosen to also be simultaneously eigenfunctions of these operators, which
leads to
1 '
F vo (x)= -ez(kSzfkZa)X vo(Y), (3.348)
27T
with k , and k , as wavevector components in the x- and z-direction. For
X,(y), equation (3.347) yields

where we have set


eB 2 liC
, yo = rCk,, r z = -. (3.350)
wcvD = Im:,lc eB
The total energy eigenvalue E , of equation (3.347) is a function Ew(k,) of
k,. It follows from the partial eigenvalue ELo of equation (3.349) associated
with motion in y-direction through the relation

(3.351)

Equation (3.349) describes a harmonic oscillator with center at yo, mass lm&1
and eigenfrequency wcw. It is well-known that in this case the eigenstates
can be described by integer quantum numbers n = 0 , 1 , 2 . . . ,00, with the
corresponding eigenenergies E L given by

EL, E &=
~ FAW,,,
( + a> . sgn(m&).
n - (3.352)

These energies are called Landau levels. For the eigenfunctions Xvo(y),we
have
454 Chapter 3. Electronic structure of semiconductor crystals with perturbations

(3.353)

with H , a Hermitian polynomial of order n. Thus the electron executes an


oscillation in y-direction with angular frequency wc, about its equilibrium
position yo. The pertinent frequency is given by

(3.354)

For mC, = 0.2 m and B = 1 Tesla, this frequency amounts to 140 G H z , i.e.
it lies in the microwave region. According to equations (3.352) and (3.353),
the spectrum of allowed energy values forms a set of 1-dimensional energy
bands, the Landau subbands,

which exhibit non-zero dispersion in the k-direction parallel to the magnetic


field.
It is enlightening to consider the density of states (DOS) of the energy
spectrum of equation (3.355), taking the conduction band vo = c as an ex-
ample. This DOS can easily be calculated by means of the general definition
of equation (2.205). One need only replace the quantum number i in (2.205)
by the set of quantum numbers n , k,, k y , and replace Ei by the set Ecn(k,)
of Landau subbands from equation (3.355). Then the DOS, p c ~ ( E )of, the
conduction band becomes, apart from a constant factor,

In Figure 3.81, p c ~ ( Eis) shown as a function of E. It becomes infinitely large


+
at the uniformly spaced energies E , = E, hLw,(n + i). Each singularity
corresponds to the minimum of a particular 1-dimensional Landau subband.
For energies slightly above E,, the DOS decays like l / d G . The inverse
square root-behavior of the DOS of a single 1-dimensional parabolic band
has already been discussed in section 2.5.
The physical meaning of the eigenstates (3.353) and eigenenergies (3.352)
of a Bloch electron in a magnetic field may be understood as follows. In
the z- and I-directions the electron propagates like a plane wave. Of the
two wavevector components k , and k,, however, only k , enters the energy
eigenvalue Euon. This unusual absence of k , is due to the fact that h k ,
represents the eigenvalue of the canonical momentum operator component
(I
455

v)
.-

% 4
Q-
2

0 I I I I 1
0 2 4 6 0
E - Eg(hwCc)

Figure 3.81: DOS in a magnetic field.

p , rather than that of the kinetic momentum operator component (pz +


:By). From the expression for yo in equation (3.350) it may be seen that k ,
determines the equilibrium position of the harmonic oscillations along the y-
axis. Since the eigenenergies Evan(kz)of these oscillations do not depend on
the equilibrium positions, a high degree of degeneracy exists with respect to
k,, and every arbitrary linear combination of eigenstates having different k,-
values is again an eigenstate. If the vector potential would have been chosen
in the form (3.320) which is symmetric with respect to I and y, then such
linear combinations would have resulted in eigenstates which also execute
oscillations with frequency wcvo in the 2-direction, but phase shifted by 7rj2
with respect to the oscillations in the y-direction. This signifies circular
motion in the plane perpendicular to the magnetic field, known as cyclot.ron
m o t i o n Such circular motion is a consequence of the axial symmetry of the
magnetic field. In classical mechanics, one also has circular orbits for the
motion of an electron in a magnetic field, projected on I - y plane.
The radius of the cyclotron orbit depends on the quantum number n,
larger values of n and, with this, also larger energies ELan mean larger cy-
clotron radii. By absorbing an energy quantum of an electromagnetic radia-
+
tion field, an electron can make a transition from the n-th to the (n l)-th
cyclotron orbit. Such transitions are involved in cyclotron resonance: At
a fixed frequency, the absorption of microwave radiation is measured as a
function of magnetic induction B. Varying the magnetic field, the cyclotron
frequency can be adjusted to match the fixed frequency of the microwave ra-
456 Chapter 3. Electronic structure of semiconductor crystals with perturbations

Magnetic Induction/ Tesla 4

Figure 3.82: Cyclotron resonance spectrum of Germanium at 4 K . (After Dexter,


Zeiger, and Lax 1956.)

diation, whereupon the absorption becomes a maximum. Using the position


of this maximum on the magnetic field scale, the effective mass can be deter-
mined. Measurement of cyclotron resonance is an important experimental
method for the determination of effective masses of electrons and holes in
semiconductors and metals. In Figure 3.82, the first peaks of the cyclotron
resonance spectrum of germanium are shown. Because of the many valley
structure of the conduction band, the electrons give rise to several cyclotron
resonance maxima. The holes show up in two well-resolved maxima in Fig-
ure 3.82, one due to heavy holes and the other due to light holes. Each of
the two hole peaks has a doublet structure which can be seen experimentally
at higher resolution (in Figure (3.82) it is absent.
To understand the cyclotron resonance spectrum of holes, the energy
eigenvalues of the rs-valence band in the presence of a magnetic field must
CJ *
be calculated from equation (3.343) with D from equation (3.74)) and H M
from equation (3.346). One finds that the energy levels can always be char-
acterized by an integer quantum number n , as in the case of the harmonic
oscillator, but in contrast to this case, each value of n determines a group of
four energy levels, two corresponding to heavy and two to light holes. B e
cause of broken time reversal symmetry in the presence of a magnetic field,
each of the doubly spin-degenerate heavy and light hole levels in the absence
of a magnetic field splits into a doublet. By means of cyclotron resonance
measurements these levels can be determined experimentally. Comparison
with the theoretical results also yields values for the constants n and q. In
Table 3.11 these values are listed for several semiconductors.
457

Chapter 4

Electron system in
t herrnodynamic equilibrium

4.1 Fundamentals of the statistical description


In the two preceding chapters we discussed the possible stationary quantum
states of the electron system of a semiconductor. We found that these states
could be represented, approximately, by Slater determinants of stationary
oneparticle states. In the case of ideal semiconductor crystals the o n e
particle states are spatially extended Bloch states, and the corresponding
energies form energy bands. For real semiconductors with shallow and deep
centers there are, in addition, localized oneparticle states associated with
energy levels in the gap between the valence and conduction bands. The
energies of extended one-particle states were found to be independent of
their occupation, i.e. independent of the specific configuration of the many-
electron system, while configuration dependencies due to Coulomb repulsion
were found for the energies of localized states. We also described calcula-
tional procedures used to determine the stationary oneparticle states and
the corresponding energy bands and localized levels, and, in limiting cases,
these energies were evaluated explicitly.
However, we have not yet addressed the question of whether the electron
system is actually in a stationary state, and if so, which of the various one-
particle states are involved in the construction of the corresponding Slater
determinant. This is equivalent to asking of which of the infinite number
of possibilities will be realized for distributing the electrons of a periodicity
region over the infinitely many one-particle states. We have yet to deal with
this question, as it is by no means already answered or even addressed by the
determination of all possible stationary states of the electron system. The
situation is similar to that in classical mechanics wherein the totality of all
possible paths of a particle system, i.e. the set of all possible solutions of the
458 Chapter 4. Electron system in thermodynamic equilibrium

mechanical equations of motion of the system, has been determined. To find


the actual path of the classical system, one needs additional information in
the form of initial conditions or initial values of the constants of the motion.
In the case of a quantum system, the initial value of its wavefunction is
required. In any case, this information would be necessary if one were seeking
for a deterministic description of the system. But this is not the case.
Macroscopic systems are generally not described in a deterministic, but a
statistical way. This means that the system is understood to be supplanted
by an ensemble of systems, all of which have the same Hamiltonian as the
original one, but may be in different quantum states. The properties of
this ensemble on average are taken to be representative of the macroscopic
properties of the system. The state of the ensemble is described by a so-
called statistical operator b. This operator has to be chosen in accordance
with the macroscopic state of the system. If the latter is a thermodynamic
equilibrium state, b may be assumed to be diagonal with respect to a ba-
sis set of stationary quantum states, in our case of Slater determinants. A
particular diagonal element of j indicates how many individual systems of
the ensemble are in the stationary state corresponding to the diagonal site
under consideration. The diagonal elements are specified using further in-
formation on the macroscopic state of the system. In our case, the electron
system may exchange energy with the atomic cores of the crystal. Further-
more, an electron transfer to and from these cores, in particular to and from
impurity atoms, may take place. In statistical mechanics, the presence of
such a heat and particle bath mandates a description in terms of the grand
canonical ensemble. This ensemble is characterized by two macroscopic state
parameters, the temperature T , and the chemical potential p. Its statistical
operator b is given by

p* = e - ( H - p f i ) / k T

where H and denote, respectively, the Hamiltonian and the total particle
number operators of the many-electron system, in the so called occupation
number representation This representation is well suited to describe Slater
determinants in a more compact way than was done in Chapter 2. Before
proceeding further, we will introduce this representation (for more detail see
Appendix C). We number the possible one-particle states by an integer i
which can take the values 1,2,. . . ,m. If a particular one-particle state i
occurs in a Slater determinant one says that it is occupied or that its oc-
cupation number Ni is 1. If the state i does not occur, then Ni is 0. A
given one-particle state cannot occur more than once, otherwise the Slater
determinant vanishes in conformance with the Pauli exclusion principle for
electrons. The occupation numbers Ni can therefore only take the values 0
and 1. Using occupation numbers, the Slater determinants can be described
4.1. Fundamentals of the statistical description 459

as vectors IN1, N 2 , . . . , N,) in the occupation number space. These vec-


tors have, altogether, as many Ni different from zero as the total number of
electrons in the system. The IN1, N z , . . . , N,) may be understood as eigen-
vectors of particle number operators i?i for the eigenvalue Ni (see Appendix
C ) . Therefore,

fiiIN1, N27... N w )
j NilN1, N 2 , . . ' 7 Nw), (4.2)
and the total particle number operator k of the electron system follows by
summing the f i i over all i,
00

N = p i . (4.3)
i=l
Moreover, (4.2) and (4.3) imply that

The occupation number vectors IN1, N z , . . . , N,) form a complete basis set
in Hilbert space of the many-electron system. These vectors may be used to
represent an arbitrary many-electron operator. If only extended oneparticle
states are involved, like in the case of an ideal crystal, the occupation number
representation of k reads
W

H= EiNi. (4.5)
i=l

where Ei are the configuration-independent oneparticle energies. If also


localized oneparticle states exist, like in the case of a real crystal having
shallow and deep centers, the energies Ei are, in general, configuration-
dependent. The Slater determinants IN1, N z , . . . , N,) are, of course, also
eigenvectors of I? so that

Using equations (4.2) to (4.6) one may easily show that the occupation
number space matrix representation of the equilibrium density matrix of
equation (4.1) is in fact diagonal with respect to the basis set of Slater
determinants "1, N z , . . . , N,), with the diagonal elements given by
460 Chapter 4. Electron system in thermodynamic equilibrium

The average value < fii > of the particle number operator fii, like the
average of any other operator of an observable quantity of the many-electron
system, is formed in accordance with the relation

Here, the symbol Tr[. .I stands for 'Trace' and means the summation over
all diagonal elements of the operator within the brackets with respect to
any basis in the Hilbert space of the many-particle system. One may easily
demonstrate that the value of the trace is independent of the particular basis
chosen. Of course, it is advantageous to use the set of Slater determinants
IN1, N z , . . . , N,) as basis, since the matrices of 6 and Ni are known in this
represent ation.

4.2 Calculation of average particle numbers


Below, we will calculate average particle numbers in equilibrium by means of
equation (4.8). In doing so it is necessary to distinguish between configuration-
independent extended one-particle states, and localized one-particle states
which do depend on the configuration of the many-electron system because
of Coulomb repulsion effects. We begin with configuration-independent one-
particle states.

4.2.1 Configuration-independent one-particle states


As above, we describe them by an integer quantum number i. Each value
of i represents only one state, with different spin orientations corresponding
to different values of i. The two trace expressions in the numerator and
denominator of equation (4.8) become, respectively,

where we have set p = l / k T for brevity. The two expressions (4.9) and (4.10)
can be written as products of sums over particle numbers for a particular
state j , as indicated in equation (4.10). In the quotient, all factors cancel
except the two with j = i. These remain and give rise to the expression
4.2. Calculation of average particle numbers 46 1

Since the particle number Ni may take only the values 0 and 1, it follows
that

(4.12)

The average particle number in the one-particle state i thus depends only
on the energy of the state. If one identifies the chemical potential p with
the Fermi energy E F , and uses the notation

(4.13)

for the Fermi distribution function previously introduced in Chapter 1, we


obtain

(4.14)
We conclude that the average particle number for configuration-independent
one-particle state i is given by the Fermi distribution, just like for an ideal
gas. The distribution function of classical statistics, the B o l t z m a n n distri-
bution

(4.15)
follows from the Fermi distribution for energies E with E - E F >> k T . This
also means f(E)<< 1. Under this condition the average particle number in
the one-particle state is small compared to 1, and thus the state is almost
unoccupied. Substantial deviations from the Boltzmann distribution occur
when E - EF is no longer large in comparison with k T and, therefore, f ( E )
no longer small compared to 1. In this case of substantial occupancy the
limited capacity of a one-particle state to host electrons (remember the Pauli
exclusion principle) becomes important, and the average particle number in
it deviates from that in classical statistics in which the hosting capacity of
states is not limited.
The deviation of the average particle number in a oneparticle state from
the value given by the Boltzmann distribution is called (statistical) degen-
eracy. A particle system is referred to as non-degenerate if it is described
by the Boltzmann distribution, otherwise it is called degenerate. The ap-
pearance of statistical degeneracy is a purely quantum mechanical effect - it
reflects the quantization of energy levels and the indistinguishability of elec-
trons. Statistical degeneracy applies to other elementary particles as well.
Indistinguishability refers to the fact that, in the exchange of two identi-
cal elementary particles, the total wavefunction of the many-particle system
transforms either into its negative (this holds in the case of electrons and
leads to the Pauli exclusion principle and with it to the Fermi statistics),
462 Chapter 4. Electron system in thermodynamic equilibrium

or the total wavefunction transforms into itself (then the population of the
oneparticle states is not limited, but the quantization of energy yields a
distribution function which also deviates from the Boltzmann distribution).
While the Fermi distribution emerges in the former case, in the latter case
it is the Bose distribution function, which may be formally obtained from
the Fermi distribution function (4.13) by setting the chemical potential zero
and replacing +lby -1in the denominator.
In the particular case of band states, the quantum number i is the pair
vk of band index v and quasi-wavevector k. In order that the wavefunction
luk) really represent only one quantum state, which is the case here, u must
also describe the spin state, or, alternatively, we must augment v to include a
spin quantum number CT. We use the latter description, i.e. we will write the
band states, henceforth, in the form lvko). With this, we assume that both
the band and spin quantum numbers I/ and u are compatible and meaningful
simultaneously. This is only justified if the spin-orbit interaction plays no
+
role and the two states (uk $) and lvk - g). have equal energy E,(k),
which we take to be the case here. In such circumstances, equation (4.14 )
yields
1
< fiuk6 > = f(E,(k)), ff = f-, (4.16)
2
and the average particle number < NUk+& + NUk-+ > for band u and
wavevector k follows as

(4.17)

The Fermi distribution (4.14) solves the problem of determining the average
equilibrium particle number < I?i > for one-particle states li) having energies
which are independent of the configuration of the many-electron system.
This distribution cannot be applied to configuration-dependent one-particle
states. In the latter case the above calculation of average particle numbers
has to repeated, taking account of the configuration dependencies of o n e
particle energies because of Coulomb repulsion effects.

4.2.2 Configuration-dependent one-particle states


We consider oneparticle states i localized at shallow or deep centers. As
in Chapter 3, a center with n electron ionization levels within the energy
gap is referred to as an n-multiply ionizable donor, and a center with n hole
ionization levels in the gap is referred to as an n-multiply ionizable acceptor.
The simplest case in which localized states occur is that of a simply ioniz-
able donor or acceptor center. However, the peculiarities of the occupancy
statistics of localized states can be understood more easily in the case of
double donors or acceptors. Therefore, we start with them. Their states
4.2. Calculation of average particle numbers 463

are described in oneparticle approximation with a configuration-dependent


Hartree potential (see section 3.2). The simply ionizable donor and acceptor
centers may be treated as limiting cases of double donors and acceptors, as
will be demonstrated below.

Double donor and acceptor centers

We assume a double donor. To be specific, we can think in terms of a


substitutional sulfur atom in a silicon crystal. The one-particle states i are
specified as follows. For simplicity, we first consider only the two 2-fold spin-
degenerate oneparticle ground states ID f ;)' = 1D)'I f $) of the 1-fold
occupied center, and the two spin-degenerate oneparticle ground states IDf
21 )2 = 1D)21f $) of the 2-fold occupied center, thereby omitting all excited
states. The pertinent configuration-dependent one-particle energy levels are
deno'ied by E b and E i , respectively. Due to the Coulomb repulsion between
+
localized electrons, we have E L = E& U where U is the Hubbard energy
of the 2-fold occupied center (see sections 3.2, 3.4 and 3.5). We assume
U > 0 so that EL > EL. The total energy of the two electrons at the 2-fold
occupied center is 2 E i - U = 2(E& - $ U ) 2 E E where E E means the
ionization level of the 2-fold occupied donor. The ground state oneparticle
wavefunctions (Da)' and of the donor center in the two occupancy
states will also differ, but the difference will be neglected here. The details
may be ascertained using first order perturbation theory with respect to
the perturbation Vfi of the Hartree potential of equation (3.18). With this
approximation, we again have only one type of one-particle wavefunctions

I D a ) l % IDu)2 = p a ) , (4.18)

as was the case before, and consequently also only one type of particle num-
ber operators Nou, u = fi,Using the pertinent occupation numbers No,
of localized states, the N-particle states of system may be characterized as
IND+rNo-+, NJ, ...Nm), where N3, ..., N, denote the occupation numbers
of ex&nded band states. We use these N-particle states as a basis for the
calculation of the average particle number < N 1 > in localized state IDa)
D2
according to expression (4.8). We have

(4.19)

The factors referring to extended states are the same in the numerator and
denominator of expression (4.16), thus they cancel. In the remaining factors
referring to localized states, the following four configurations are possible:
No; = 0, N0-i = 0, N o + = 1, ND-l= 0, No+ = 0, ND-1= 1, No+ =
2
464 Chapter 4. Electron system in thermodynamic equilibrium

1, N D - ~= 1. Considering these configurations and their energies specified


above, the evaluation of expression (4.16) results in

1
[,-P(%-P) + .-2P(E2d-d] , 2
= f-. (4.20)

The average number of particles occupying the center is the sum of < I$ 1 >
D;i
and < iD-+
>, given by

Using this expression, we will now discuss how the average number of par-
ticles localized at a center is influenced by configuration dependencies of
one-particle energies due to Coulomb repulsion. In Figure 4.la this num-
ber is shown as a function of chemical potential p for T = 0 K . As long
+
as p < EA, we have < I? 1 f i D - ~> = 0. If p exceeds EL, then
D;i
< N
D5
1 + kD-&> suddenly increases to the value 1, which is maintained
as long as E b < p < E g holds. When p becomes larger than E g , then
+
< fi 1 ND-+ > increases to the value 2, which is never exceeded even if
D-i
p increases further. In regard to dependence on chemical potential one has,
therefore, the values 0, 1 or 2 for the average particle number at the center.
If there were no Coulomb repulsion effects, then E g = E& would hold and
the 1-fold occupied state would cease to play a distinct role, as in the case
of band states (see Figure 4.lb). For temperatures above zero Kelvin, the
+
variation of < N 1 kD-i> with p is qualitatively the same, except that
D2
the sudden changes which occur at zero temperature are smoothed out. The
fact that the ionization level E$ rather than the oneparticle level E$ of the
2-fold occupied donor center determines the probability for two electrons to
be at the center, is physically understandable: on a microscopic scale, ther-
modynamic equilibrium is a dynamic state in which ionization (excitation)
and de-excitation processes balance each other.
The above results for a double donor may be readily transferred to a
double acceptor. If one replaces D by A , equation (4.21) gives the average
number < $ 1
AT
+ > of electrons (rather than holes) at such a center,
provided the energy levels EA and Eil in this expression are understood
as electron rather than hole energies. Thus the average number of holes is
(2- < NA+ 4 NA-+ >).
The method of calculating the average particle number at double donor
and acceptor centers may be easily generalized to centers with more than
4.2. Calculation of average particle numbers 465

. ElD E':
1
ED= ED
21
E:
21
ED+w

+
Figure 4.1: Average electron number < N D 3 f i D - + > according to formula
(4.21) of a double donor for T ---t 0 K as a function of chemical potential p. The
energy levels E A and E$ are both negative but different from each other in case
a), both negative and equal in case b), and only E A is negative in case c) while E S
tends to +m. In the latter case one actually has a simply ionizable donor. In case
b) there are effectively no configuration dependencies due to Coulomb repulsion
which is actually only true for band states.

two ionization levels in the energy gap. For n-fold ionizable centers one has
+
n 1 charge states. In calculating the average number of particles at such
centers one must include all of them, and take account of the fact that their
oneparticle energies and ionization levels differ. We omit this calculation
because it is rather cumbersome, and does not provide new physical insight.
Instead we turn to simply ionizabIe centers which, from a practical point of
view, are particularly important ones in semiconductors.

Simply ionizable donor and acceptor centers

Formally, the transition from a double donor to a simply ionizable donor can
be represented by E E becoming infinitely large. In this case, E b plays the
role of the binding energy E D of the simply ionizable donor, and all terms
in equation (4.21) involving E g automatically vanish. Thus one obtains

which may be written in the form

(4.23)
466 Chapter 4. Electron system in thermodynamic equilibrium

with
(4.24)

as an effective distribution function. The latter can be interpreted as the


probability that the simply ionizable donor is occupied by 1 electron. This
carries the implication that f * ( E )has the same meaning for localized states
of simply ionizable donor centers as f(E)has for the doubly spin-degenerate
band states.
Alternatively, equation (4.22) can be obtained as follows. One again
calculates the average value < N 1 +
D%
kD-;> according to the general
expression (4.8), but only takes into account the three states No; = 0 and
N D - l = 0; N 1 = 1 and N D - ~= 0; as well as N D f = 0 and No-+ = 1.
Dg.
This procedure, starting from equation (4.8), yields expression (4.22).
It is appropriate to clarify the consequences of Coulomb repulsion effects
here. In the limiting case of strong degeneracy, the exponential term in
the denominator of expression (4.24) for f*(Eo)can be neglected, and one
arrives at ED) = 1 (see Figure 4 . 1 ~ ) . For band states, one obtains,
under the same conditions, according to equation (4.19), the average value
+
< Nyik fiv-tk > = 2. Therefore, the energy level localized at the
donor can host only half of the particles which can be accommodated by
a band level. If statistical degeneracy exists, i.e. if the localized levels are
almost completely occupied, then Coulomb repulsion limits the occupancy
capacity of the localized states. In the non-degenerate limiting case, on
the other hand, in which the 1 in the denominators can be neglected and
+
< N o + No-4 > << 1 holds, expression (4.24) yields the Boltzmann
distribution multiplied by 2. The same average particle number is obtained
in this limiting case for a doubly spin-degenerate band state. This result
is understandable, because in the non-degenerate case most of the localized
levels are empty or simply occupied, and Coulomb repulsion does not play
a significant role.
The results derived above for simply ionizable donor centers are imme
diately applicable to simply ionizable acceptors. If one replaces D by A in
equation (4.23) , then

becomes the average number of electrons at the acceptor. The average hole
number at the acceptor follows from this as

(4.26)

Here, the one-electron energy E A of the acceptor state represents the binding
energy E B of the hole at this level.
4.2. Calculation of average particle numbers 467

For a strongly degenerate hole distribution, i.e. for E A - p >> k T , the


exponential term in the denominator of expression (4.26) may be neglected,
and one gets 1 - ~ * ( E AM ) 1. This is half of the number of holes which under
the same conditions, would be hosted by a valence band state. If the hole
distribution is non-degenerate, expression (4.26) becomes the Boltzmann
distribution function, multiplied by f. The latter factor reflects the fact
that the population of the acceptor level by electrons is almost doubled due
to the 2-fold spin degeneracy of the acceptor ground state.
Thus far, the average particle numbers at localization centers have been
calculated under the assumption that, beside the spin degenerate ground
state, no other energy levels exist with wavefunctions localized at the center.
As we know from Chapter 3, there are, however, generally, several such
states, indeed, infinitely many at hydrogen-like donors and acceptors. We
refer to them as excited localized states. Below, we will determine the average
particle numbers at centers having such excited localized states.

Centers with excited localized states

We restrict our considerations here to simply ionizable centers, although


there are also excited states for multiply ionizable centers. The roles of these
states are similar for both kinds of centers, and we will discuss the simpler
case here. We examine a hydrogen-like donor, for which the localized states
are characterized by the three orbital quantum numbers n, 1 , and m , and the
spin-quantum number c. We denote the wavefunctions by IDnlmu), and the
corresponding energy levels by ED,. The degree of degeneracy of level n is
denoted by gDn. For the hydrogen atom, g D n = 2n2. w e seek the average
number of electrons at the center, which is just the average particle number
in the space of all states IDnlmc), < C n l m o N ~ d m o >. The averaging
involved has to be carried out over states of the N-electron system in which
either no particle is found at the center, i.e. with CdmU N I ) , =~ ~0, ~or
there is exactly one, i.e. with Cnlmu ND~~~ , Then, according to the
=, 1.
general formula (4.8) for the average value, we obtain

< C fiDnl-
nlmu
> = 1 + C g D n e -P(ED,-P)
[ n I-' n
gDne-fl(EDn-p),

(4.27)
This expression may be written in the form (4.24) if the ground state energy
ED' is denoted by E D , and if the factor f multiplying the exponential term
in the denominator in (4.24) is replaced by the temperature-dependent factor

(4.28)
468 Chapter 4 . Electron system in thermodynamic equilibrium

With this we have

(4.29)

The validity of this rewriting may be easily demonstrated by inserting y*


of equation (4.28) into equation (4.29). Omitting the excited states in y*,
considering therefore only the term with n = 1 and g D 1 = 2, we obtain the
4
value for y*,and the average particle number at the donor center is given
by equation (4.24). Considering the excited states, 7*becomes smaller than
4, provided the temperature T is larger than 0 K . At T = 0 K , y* retains
4
the value even in the presence of excited states. This value decreases only
slightly with increasing temperature, as long as kT << [ED, - ED]holds for
all excited states n 2 2. This inequality provides a simple criterion which
helps decide whether or not the excited states are essential in the calculation
of the average particle number at the center. The formal evaluation of the
factor y* for hydrogen-like states with E D , = E, - E g / n 2 and g D n = 2n2
at arbitrary temperature T > 0 K leads to the result y* = 0, so that an
apparently strong influence of the excited states occurs at all temperatures
T > 0 K . In reality, however, the orbital radii U D of ~ the excited states
are limited, if not through other factors such as finite sample size, then
by the adjacent donor atoms - their distances from each other must be
large compared to 2 U D n , so that the n - t h electron shell may be formed
unhindered by the adjacent donors. For a donor concentration of 1015 ern-'
this condition yields 2 u <<~1000~ A. In silicon, 2UDn equals about 100 A x n ,
so that n must clearly be smaller than 10 and the overall influence of the
excited states for T > 0 K is not so very great.
The factor y* significantly influences the average particle number at a
donor center only for weak occupation of the center, which occurs when
y* exp[P(ED - p ) ] >> 1 or E D- p >> kT holds, whence equation (4.29) yields

4,
Since y* < the average particle number at the center is enlarged as com-
pared to the case without excited states where y* = 4.
The excited states
therefore lead to an increase of the average particle number at the cen-
ter. This is physically understandable because more states are now avail-
able for occupancy than without excited states. If, on the other hand,
p - E D >> kT, i.e. the ground state is almost completely occupied, then
one has y*exp[P(ED - p ) ] << 1, and correspondingly

(4.31)
4.3. Density of states 469

so that one gets the same average particle number at the donor center as
without excited states. This is also physically clear immediately - because
the donor can host only 1 electron, and, with high probability it will occupy
the ground state, so excited states play practically no role.
As in the case without excited states, the results for donors can also be
transferred directly to acceptors but we will not write down the correspond-
ing expressions explicitly.

4.3 Density of states


Having determined the average particle numbers in single one-particle states
and sets of such states, we can now proceed to explore the distribution of
the electrons of a semiconductor over these states.

4.3.1 Total electron concentration


The total electron system is composed of the valence electrons of all atoms
of the semiconductor. We denote their number by N b t a l . This number
determines the average total particle number < fi > of the system,

<N >= Nbtal. (4.32)


This relation may also be understood as a condition which determines the
chemical potential p in the statistical operator j which is used in the aver-
aging procedure. Considering equation (4.3) for N as a sum of one-particle
number operators f i i for the various one-particle states li), one obtains from
equation (4.32)

(4.33)
i= I
i)
The one-particle states l i ) comprise band states l v k f and localized states
lyf i)where y abbreviates the orbital state quantum numbers and the posi-
tion of the center. The band states are assumed to be 2-fold spin-degenerate,
as before. With this, we presuppose that spin-orbit splitting is negligibly
small which, as we know, is often not justified. The generalization of the
results to the case in which spin-orbit splitting has to be taken into account
i)
is straight forward. The localized states (7 f are taken to be the ground
states of simply ionizable donors (y = D ) or acceptors (y = A ) at various
positions in the crystal. We assume that there are no excited localized states.
The i-sum of equation (4.33) is decomposed in two partial sums, one over
the band states, and one over the localized states. Correlations between the
two sets of states which disturb this decomposition are neglected. With this,
we obtain the expression
470 Chapter 4 . Electron system in thermodynamic equilibrium

For the band part, equation (4.17) yields

(4.35)

with f(E)as the Fermi distribution function. The localized part in equation
+
(4.34) can be written as a sum over the average values < k 1 k7-; > for
7F
the various centers 7,because they are independent of each other, whence,

Using the effective distribution function f*(E)for localized states from equa-
tions (4.24) and (4.25) we have

Altogether, we obtain

NtOta1= +
2 C f ( E u ( k ) ) Cf*(E,). (4.38)
uk 7
The total electron number Ntotd is an extensive quantity. Generally, the
corresponding intensive quantity, i.e. the total electron concentration related
to the volume of the periodicity region, is of interest. It is defined by the
expression

Ntotal
nbtal = -
0 '
(4.39)

In the further evaluation of the relation (4.38) we employ nthl instead of


Ntotal.

4.3.2 Density of states of ideal semiconductors

We first consider equations (4.38) and (4.39) in the absence of localized


states, i.e. for an ideal semiconductor. Localized states will be taken into
account in the next subsection dealing with real semiconductors. For the
total concentration ni%$ of electrons occupying the extended Bloch states
of an ideal semiconductor one obtains, using (4.38) and (4.39), the expression
4.3. Density of states 471

(4.40)

Introducing a &function we rewrite the sum on k in the following way:

(4.41)

Substituting this expression in equation (4.40) for n i i l yields

(4.42)

where

2
pid"'(E) = - C 6 ( E - E , ( k ) ) . (4.43)
uk

is the density of states (DOS) of the ideal semiconductor considered in Chap-


ter 2 (see equation (2.210). The introduction of the density of states in ex-
pression (4.42) for nh%i is significant in that it provides a physically mean-
ingful division of the information needed for the calculation of the average
total particle number and other statistical average values into two parts:
first, information about the possible stationary quantum states of the sys-
tem is contained in the density of states pidea'(E), and second, information
about the statistical occupation of the states of the system is reflected by
the distribution function f(E).One may also say that p*'(E) contains
information relating to the dynamics of the system (classically, this would
pertain to the solutions of Newton's equation of motion), and f(E)informa-
tion about the statistics of the system (classically this would reflect on the
statistical distribution of initial conditions or constants of motion). These
two types of information are largely independent of each other and both are
needed to calculate statistical average values in thermodynamic equilibrium.

Isotropic parabolic two band model

We consider a model of a semiconductor which has only one valence band


and one conduction band. Both bands are assumed to be parabolic and
isotropic for energies close to the band bottoms. The DOS pikd(E) of the
conduction band is then given by equation (2.220), with v identified as the
conduction band index c, and E d as the gap energy E,, so that

ideal I 2m* 3/2


pc ( E )= -
2x2 (L)
h2 B(E - E , ) d G . (4.44)
472 Chapter 4. Electron system in thermodynamic equilibrium

Figure 4.2: Density of states of an ideal


E
semiconductor with one conduction and
one valence band. The full line corresponds
to a parabolic dispersion law, the dashed
line is closer to reality.

1 ENERGY
GAP

The DOS of the conduction band thus exhibits square root-like behavior at
the band edge (see top part of Figure 4.2). Formally, it continues to follow
a square root law also for energies deep in the band. However, for such
energies, expression (4.44) does not apply. Qualitatively, the shape shown
by the dashed curve in the top part of Figure 4.2 is more realistic.
Expression (2.220) may not be used directly for the valence band, since its
derivation presupposes that the effective mass m t is positive, while for the
valence band it is in fact negative. The calculation above can, however, be
modified easily to the case m: < 0. With v = v, E,o = 0, and mz -mt >
0, we have

(4.45)

The density of states of the valence band is thus seen to have square root-like
dependence at the upper band edge and formally also continues to increase
according to a square-root law for energies deeper in the band (see bottom
part of Figure 4.2). The dashed curve again shows the more realistic shape.
The entire density of states pideaz(E)of an ideal semiconductor having one
valence and one conduction band is given by the expression

In the energy gap the density of states pideal(E)vanishes, as a direct con-


4.3. Density of states 473

sequence of the fact that no stationary electron states exist there (see Fig-
ure 4.2). Finally, we write down the total electron concentration n&%: of
the ideal semiconductor having one conduction band and one valence band.
From equation (4.42) it follows that

Bands with more complex structures

In the case of the elemental semiconductors Si and Ge, the conduction band
edges consist of several valleys, and the iso-energy surfaces of the valleys
are anisotropic. This has consequences for p,(E). For Si, for example, the
conduction band density of states p p ( E ) takes the form

The valence band edges of diamond and zincblende type materials are de-
generate. The degree of degeneracy depends on whether spin-orbit splitting
is important or not. In the case of Si, the spin-orbit splitting energy A is
small in absolute units (44 m e V ) , but still large enough to compete with k T
(25 meV for T = 300 K ) . Thus, for temperatures that are not extremely
high, most of the holes are hosted by the I?$ valence band while the spin-
split I?$ band remains practically devoid of holes. This means that there are
two hole states per k-vector (apart from the 2-fold spin degeneracy), one for
heavy holes, and one for light holes. Thus , the corresponding hole density
of states p p of Si, calculated in the isotropic parabolic approximation of
equation (4.45), is given by the expression

The same expression applies for Ge and GaAs.


In cases where non-parabolicity effects are essential, the energy depen-
dence of the density of states no longer follows a square-root law. This is
surely true for energies far removed from the critical point. In narrow gap
semiconductors this occurs, however, even close to the critical point. For
example, as shown in section 2.7, (see equation (2.404), the electrons and
light holes in semiconductors having small energy gaps are, within the Kane
model, described by the dispersion relation
4 74 Chapter 4. Electron system in thermodynamic equilibrium

(4.50)

The conduction band density of states p P ( E ) can be derived from equa-


tion (4.50) as follows: First of all, the general relation

p p ( E )= 4
27r
k 2 S ( E - E,(k)) (4.51)

is still applicable. Here, we introduce E, as new integration variable, with the


function k ( E c ) taken from the dispersion law (4.50). In this way, we obtain
p p ( E ) for E-values above the conduction band edge at Ec = E , +4 as
follows:

Pc
Kane(E) 1
-
27r2
(-)3
2P2
3/2 (.- 5 - 3)
3 2
.-/
(4.52)
If E lies close to E,, then p p ( E ) given by (4.52) takes the form of ex-
pression (4.44) for a parabolic band. To see this, the effective mass rn; has
to be replaced by the expression from Table 2.13. That is to be expected,
because the band structure is also almost parabolic for energies close to the
edge within the Kane model.

4.3.3 Density of states of real semiconductors

We now consider doping the ideal semiconductor discussed above with N D


donor atoms and N A acceptor atoms per cm3 to obtain a keal semiconduc-
tor. Each donor atom is assumed to give rise to a 1-fold ionizable donor
level E D , and each acceptor atom gives rise to a 1-fold ionizable acceptor
level EA. Excited donor or acceptor states are omitted from this considera-
tion. The assignment of the donor and acceptor levels to impurity atoms is
not essential in what follows. In principle, these levels can be generated by
any simply ionizable point perturbation, including structural defects, with
the same results. This also implies that the levels need not be shallow, they
may be deep, although the shallow case is more typical than the deep case
for the considerations below.
The electron system of a semiconductor with simply ionizable donors
and acceptors may be thought to arise from that of the corresponding ideal
semiconductor with one electron added per donor atom, and one electron
is removed per acceptor atom. The electron concentration nL7; of the real
semiconductor is thus given by the relation
4.3. Density of states 475

The concentration (1/0)&=D,A < k 1 kT-+


72
+ > of electrons occupying
the donor and acceptor levels may be calculated by means of equations (4.23),
(4.24), and (4.25), with the result

As in the context of the total electron concentration (4.42) of an ideal semi-


conductor, we introduce here the density of states pTal(E) of the real semi-
conductor. The contribution arising in the energy gap is denoted by p E $ ( E )
and we set

For this relation to correspond to equation (4.54), pi$ must be defined by


the expression

Teal ( E ) = NDG(E - E D )
pgap + N A ~ ( E- E A ) . (4.56)

The densities of states of the valence and conduction bands of the real semi-
conductor will be denoted by pLd(E) and p:"'(E), respectively. The band
states of real semiconductor are, of course, no longer pure Bloch states, but
Bloch states perturbed by scattering from the impurity atoms. Similar to
pure Bloch states, they are, however, still spread out over the entire crys-
tal. Correlation effects therefore play no important role in their occupancy.
This means that the total electron concentration in the bands of the real
semiconductor may be expressed by p;T'(E) in the same way as the total
electron concentration of the ideal semiconductor was expressed by p Z d ( E )
in equation (4.47). Therefore,

,Teal
total = 1,
0
dE P:az(E) f(E) + /Or d E PF%)
Eg
f(-w
00
+I,dE PEW f*(E) . (4.57)

The band densities of states pLY'(E) of the real semiconductor differ from
p ; Y ' ( E ) in accordance with the considerations of sections 3.4 and 3.5, where
it was shown that localized states occur at the expense of the band states
476 Chapter 4. Electron system in thermodynamic equilibrium

(Levinson's theorem). The number of conduction band states therefore de-


creases by the number of donor states, and the number of valence band states
decreases by the number of acceptor states, related to the volume unit in
each case. Thus we have

0
dE pLal(E) = dE p i k d ( E ) - NA, (4.58)
--M

(4.59)

The entire density of states

of the real semiconductor satisfies the conservation law of the density of


states, such that

l,m
dEp'"l(E) = Jrn
-m
dEpiM(E). (4.61)

Concerning the band state densities p,':'(E), we initially know, apart from
the integral relations (4.58) and (4.59), nothing further than that it must
reproduce p Z z ( E ) if N D and N A are zero. Below, we will see that it is
possible, under certain conditions, to replace the real state densities p T f ( E )
approximately by the ideal state densities p : p z ( E ) even when N D # 0 and
N A f 0.
In the following, we calculate the average electron concentrations in the
energy bands and the donor and acceptor levels separately. Such separate
calculations of the concentrations are necessary because the electrons in these
different parts of the energy spectrum behave quite differently. The electrons
localized at an impurity atom, for example, do not contribute to electric
charge transport. The electrons of the conduction band do, just as do the
electrons of the valence band, the latter, however, provide such a contribution
in a different sense, not as electrons, but as holes. Therefore, facilitate to
calculation of the conductivity of a semiconductor, one needs the average
electron concentrations in the conduction band, in the donor and acceptor
levels, and in the valence band, separately. The same remarks apply to the
calculation of other measurable quantities.
4.4. fiee carrier concentrations 477

4.4 Free carrier concentrations


We start with general considerations which refer to both ideal as well as real
semiconductors. The object is to derive a condition which fixes the total
number of electrons of these systems.

4.4.1 Conservation of total electron number


Consider (4.53) for the total electron concentration nL%t of a real semicon-
ductor. In this relation, the total concentration n&%! of the ideal semicon-
ductor may be taken from equation (4.47), which we write down for T = 0 K .
Since the Wrmi level of an ideal semiconductor lies in the energy gap ac-
cording to the considerations of Chapters 1, one has f(E)= O(EF - E ) for
T = 0 K . With this, relation (4.47) takes the form
0
nzdeal
total - S _ _ d E P"Vd"YE). (4.62)

This equation says that the total concentration nitg: of electrons equals
the total concentration of valence band states of the ideal semiconductor.
This statement is, of course, independent of temperature. The total electron
concentration of the real semiconductor may therefore be written for any
temperature as
0
n Ttotal
fd -
1, d E pid"'(E) + N o - NA. (4.63)

Employing equation (4.58) this yields

(4.64)

On the other hand, we have relation (4.57) for


p z z ( E ) using (4.56) to obtain
nTzL, where we replace

(4.65)

From this equation we subtract equation (4.64), h d i n g

(4.66)
478 Chapter 4. Electron system in thermodynamic equilibrium

Consider the abbreviations

(4.67)

These quantities have direct physical meaning: n is the concentration of


electrons in the conduction band, p the concentration of holes in the valence
band, Ngf is the concentration of non-occupied (ionized) and therefore simply
positively charged donor atoms, and NX the concentration of occupied and
thus simply negatively charged acceptors. The electrons in the conduction
band and the holes in the valence band are freely mobile charge carriers, as
pointed out in Chapter 1. In these terms, equation (4.66) becomes

n- p - ND+ + N; = 0. (4.71)

This expresses the conservation of total electron number for a real semicon-
ductor. The electrons may be redistributed between the four energy regions,
i.e. valence band, acceptor levels, donor levels and conduction band, but
their total number cannot change thereby. One may also understand rela-
tion (4.71) as the neutrality condition for a real semiconductor - the negative
charge of the electrons of the conduction band and of the ionized acceptors
must compensate for the positive charge of the holes of the valence band and
the ionized donors.

4.4.2 Free carrier concentration dependence on Fermi en-


ergy. Law of mass action.
We now derive an explicit relation between the concentrations n , p of elec-
trons and holes in the bands, as they are given by equations (4.67) and
(4.68), and the Fermi energy EF. The Fermi energy, in turn, is determined
by the neutrality condition (4.71). In the non-degenerate case of an ex-
ponential Boltzmann distribution for f(E), the position of the Fermi level
has no influence on the relative distribution of the electron occupancy over
the various available energy levels, only the total electron concentration de-
pends on EF. However for a degenerate electron gas, i.e. in the regime of
quantum statistics, the Fermi distribution function f ( E ) mandates that the
relative distribution of electrons over the energy levels is also determined by
4.4. nee carrier concentrations 479

the Fermi energy. In particular, this means that the relative distribution
depends on the total particle concentration. By changing the total concen-
tration, the ratios of the average particle numbers in the various energy levels
can be altered. This is a clear manifestation of the statistical correlation be-
tween the electrons of an ideal Fermi gas. It also shows that the chemical
potential, or Fermi energy, is much more important in quantum statistics
than in classical statistics.
In evaluating the integrals of (4.67) and (4.68), we note that the occu-
pancy factors f (E) and [l- f ( E ) ]decrease monotonically with, respectively,
increasing or decreasing energy and effectively cut off the integration range.
The cutoff energies are separated by an interval of about kT from the respec-
tive band edges. We assume that the densities of states of a real semicon-
ductor, which are required in equations (4.67) and (4.68) for the effectively
contributing energy intervals, may be approximately replaced by the state
densities of the ideal semiconductor. This approximation is justified as long
as the concentrations of donors and acceptors is small compared to the con-
centration of significantly contributing band states. Furthermore, we assume
that in the contributing energy intervals, the densities of states (4.44) and
(4.45) of the ideal semiconductor, which are predicated on parabolic band
structure, are approximately valid. Thus we obtain

(4.72)

(4.73)

Defining the Fermi integral as

(4.74)

the relations (4.72) and (4.73) may be written as

(4.75)

(4.76)

with

(4.77)
480 Chapter 4. Electron system in thermodynamic equilibrium.

In the frequently occurring case that the Fermi level lies not only in the
energy gap, but is separated by at least several k T from the two band edges,
the arguments of the Fermi integrals in (4.75), (4.76) are negative and have
absolute values large compared to 1. For such arguments the Fermi integral
is approximately given by

(4.78)
so that

n = Nce(EF-Eg)/kT, ( E , - E F ) >> k T , (4.79)

p = NVevEFfkT , E F >> k T . (4.80)


The same result for n would be obtained if the Fermi distribution f(E)for
the electrons of the conduction band were approximated by the Boltzmann
distribution (4.13) from the very beginning, as is possible for ( E - E F ) >> k T .
An analogous approximation for the electrons of the valence band is not
available, since there E < EF holds. One may apply this approximation,
however, to the holes of the valence band, i.e., to
1
- f(E)= e(E~-E)/rCT+ 1 ' (4.81)

For ( E F - E ) / k T >> 1, we have, approximately,

(4.82)
As in the case of conduction band electrons, the holes of the valence band are
also described by a Boltzmann distribution if the Fermi energy is sufficiently
remote from the band edges. One may say in this situation that the electron
system of the conduction band and the hole system of the valence band are
non-degenerate carrier gases. We emphasize that in the case of the valence
band this statement refers to the holes and not to the electrons, which are
highly degenerate. In further considerations we will always assume that the
electrons of the conduction band and the holes of the valence band form
non-degenerate carrier systems. The expressions (4.75) and (4.76) for n and
p , respectively, provide a simple criterion for the validity of this assumption:
The electrons are non-degenerate if n << Nc and the holes are non-degenerate
if p << Nu. This may be seen as follows. If n << Nc holds, then the Fermi
integral F L [ ( E F- E , ) / k T ] in (4.75) must be much smaller than 1, which
is possible only if - ( E F - E , ) / k T >> 1, meaning that degeneracy does not
exist. The same argument holds for holes.
The quantities Nc and Nv in equation (4.77) have the dimensions of spa-
tialdensities and are termed the effective densities ofstates of the conduction
4.4. f i e e carrier concentrations 481

and valence bands, respectively. This terminology is clarified by examining


relations (4.72) and (4.73) for n and p using Boltzmann distribution func-
tions for both electrons and holes. In the corresponding expressions, the
densities of states are summed over all energies, multiplied by weighting fac-
tors which account for the occupancy of the various energy states, that is,
for their effective contributions to average particle numbers. Equation (1.24)
for Nc in Chapter 1 also demonstrates this explicitly (the symbol pc of (1.24)
for a pure semiconductor corresponds to p:& here).
We now estimate the order of magnitude of the effective densities of states.
Generally, we have

Evaluating this with T M 300 K and m:,, FZ 0.5 m , then Nc,, M 10


follows. This order of magnitude may also be obtained in a strikingly simple
way: The ordinary density of states of a band v is given by equation (4.43)
as

(4.84)

which is of the order of magnitude of the number of primitive unit cells per
cm3, i.e. M ~ m - ~ With
. this number, the order of magnitude of the
effective band densities of states is obtained by multiplying it by the ratio
of k T to the whole band width. With a value of k T M 10 meV and a band
width of 10 e V , N,,, FZ lo1 follows.
The values of n and p determined by equations (4.75), (4.76) still depend
on the Fermi energy E F , which is, as yet, unknown. The equation which
determines E F is the neutrality condition (4.71). In this equation all quan-
tities may be considered to be known except the Fermi energy, for which the
equation is to be solved. Even without any calculation it is clear that the
position of the Fermi level will depend on the donor and acceptor concentra-
tions N D and N A . This dependence is transferred directly to n and p. There
is, however, a simple function of the two concentrations, their product np,
which in the case of non-degenerate carrier systems does not depend on E F ,
and hence not on N D and N A . This may be seen using equations (4.79) and
(4.80), from which it follows that

np = ni,
2 (4.85)
with
ni = (NcNV)3e-iEg/ICT. (4.86)
482 Chapter 4. Electron system in thermodynamic equilibrium

Table 4.1: Intrinsic carrier concentrations ni at T = 300 K (in ~ m - ~ ) .

GaP GaAs Si Ge Ids InSb

n, 2.7 x 10' 1.8 x lo6 1.5 x 10" 2.4 x 1013 8.6 x lOI4 1.6 x 10l6

This equation may be understood as the law of mass action for a 'chemical
reaction' in which bound electrons in the valence band and at the donor
atoms on the one hand, and bound holes in the conduction and at the ac-
ceptor atoms on the other hand, generate free electrons in the conduction
band and free holes in the valence band. The concentration of the bound
electrons can approximately be set equal to the effective density of states
Nu of the valence band, since in the non-degenerate case, the concentra-
tions of holes and of occupied donors are small in comparison to N,,. A
similar statement holds for the bound holes which, here, also include the
non-occupied states of the conduction band. Their concentration roughly
equals the effective density of states Nc of the conduction band. According
to the mass action law, one then has, for an ideal gas, n p = K ( T ) N , N , , with
K ( T ) = exp[-(gc +
g,,)/,kT]as mass action constant. Here, gc and g,, are,
respectively, the free enthalpies of an electron in the conduction band and
+
of a hole in the valence band. With gc gv = E,, equation (4.86) is indeed
the mass action law, as stated above.

4.4.3 Intrinsic semiconductors


We consider the particular case of an ideal semiconductor, such that N D
and N A both vanish. Then N & and N A are also zero, and the neutrality
condition (4.71) becomes

n - p = 0. (4.87)
This equation reflects the fact that the electrons of the conduction band
are created solely by thermal excitation of electrons from the valence band,
leaving behind an equal number of holes. If we substitute p = n in equation
(4.85), then we have

n = ni, p = ni. (4.88)

This relation was already stated in Chapter 1 on the basis of empirical facts
(see equation 1.20). One refers to ni as intrinsic concentration, because it
4.4. Free carrier concentrations 483

depends, aside from temperature, only on quantities which are determined


by the ideal semiconductor itself. In this context, the latter is called an
intrinsic semiconductor. For the intrinsic carrier concentration, with Nc =
~ ,, = 1 eV,T = 300 K , we obtain the order of magnitude
N , = lo1' ~ r n - E
ni --" 10" ~ r n - ~ Table
. 4.1 gives an overview of the ni-values for some
important semiconductors at room temperature. In this, there is a strikingly
large variation between a few carriers per crn3 for the wide gap semiconductor
Gap, and 10l6 carriers per cm3 for the narrow gap semiconductor InSb. In
Figure 4.3 the variation of ni with temperature is shown. Evaluating all
numerics and fundamental constants, we find

The position of the Fermi level E F ~for a non-degenerate intrinsic semicon-


ductor follows from equations (4.79) and (4.86) by taking the logarithm. The
result is

1 1
EF=
~ -E,+ -kTh (4.90)
2 2

Replacing Nc and N , here by equation (4.77), it follows that

1 3
EF=
~ -Eg
2
+ -kT
4
h (4.91)

For T = 0 K , the Fermi level thus lies exactly in the middle of the gap
between the valence and conduction bands. It remains there even for finite
temperature if rn; = rn:. If rn; # m:, a temperature dependence emerges
which is schematically shown in Figure 4.4. For rn: > rn; or N , > N,, EF(
moves away from the valence band edge, and for rn; > rn; or Nc > N,,
away from the conduction band edge. This may be understood as follows:
If the effective density of states in the valence band is larger than that in
the conduction band, one would have more holes in the valence band than
electrons in the conduction band if the Fermi level were to stay in the middle
of the gap. It must move away from the valence band edge to assure the
equality of n and p . An analogous interpretation can be given in the case of
Nc > N,. In short, one can say that the Fermi level is repelled by regions of
high density of states, and is attracted by regions of low density of states.
We next determine n and p for the case of real semiconductors with
donor and acceptor atoms present. In Chapter 1, the designation eztrznsic
semiconductors was already introduced to describe them.
484 Chapter 4 . Electron system in thermodynamic equilibrium

Figure 4.3: Intrinsic concen- &T/OC


trations of several semicon-
ductors as functions of tem-
perature (calculated values).

IO~/T/K- -.
Figure 4.4 Variation of
Fermi level with tempera-
ture for an intrinsic semi-
conductor.

4.4.4 Extrinsic semiconductors

n-type-semiconduct ors

Initially, we assume that there are only donors, and no acceptors are present,
N D # 0 and N A = 0. The neutrality condition (4.71) is then transformed
into an equation for the electron concentration n. In this regard, we need
the concentration Ngf = N D ( - ~ ED)] of ionized donors. The donor
population probability f * ( E D ) is given by equation (4.24) as
4.4. fiee carrier concentrations 485

(4.92)

where EB denotes the donor binding energy. In general, the donor levels,
unlike band energies, do not lie well above the Fermi level. They may be
close to, or even below it. This means that the degeneracy of the Fermi
distribution, which could be neglected for carriers in the bands, must in
general be taken into account in the effective donor distribution function
!*(ED). The expression (4.92) for f*(Eo)must therefore be used in its
original form. The Fermi energy can be eliminated from it by introducing
the electron concentration n given in equation (4.79), with the result

(4.93)

(4.94)

Apart from the factor 1/2, the quantity nl represents the electron concen-
tration in the conduction band of a non-degenerate ideal semiconductor with
a (fictitious) Fermi level at the position of the donor level. If we replace p
in terms of nT/n using equation (4.85), then the neutrality condition takes
the form

(4.95)

From this equation one may easily derive the magnitude relations

(4.96)
In a semiconductor containing donors but no acceptors, the electron concen-
tration is always larger than the intrinsic concentration ni, and the latter is
larger than the hole concentration p. It is therefore called an n-type semi-
conductor. The electrons are majority carriers in this case, and the holes
are minority carriers. To solve the neutrality condition (4.95) for n, we
transform it into the third-order polynomial equation

n3 + nln2 - (ni2 + n l N D ) n - ninl


2
= 0. (4.97)
The solution n of equation (4.97) can be obtained in closed analytical form.
The minority carrier concentration p follows from n by means of the relation
n p = nz. We examine here an approximate solution of equation (4.97) in two
typical limiting cases: 1) the case of extremely small donor concentrations
N o , in which N D << ni holds, and 2) the case of large donor concentrations
with N o >> ni.
486 Chapter 4. Electron system in thermodynamic equilibrium

1) N D <( ni :
Considering the inequalities (4.96), the term proportional to N D in equation
(4.97) may be neglected in comparison with the second term. Thus we have,
approximately,

n3 + n1n2 - n f n - ninl
2 = 0. (4.98)
The only positive real solution of this equation is

n = ni. (4.99)
For p , one gets p = p i . This means that in case l),the carrier concentrations
n and p do not differ from those of an intrinsic semiconductor. If the values of
N D are not extremely large, then the inequality N D <( ni may always be ful-
filled, simply by making the temperature and therefore also ni large enough.
For sufficiently high temperatures and not overly large N D values, one thus
always has the intrinsic case. Only if the donor concentration becomes suffi-
ciently large, and the temperature so low that the inequality N D << ni does
not hold, then the donors can influence the carrier concentrations n and p
considerably.
We turn now to the other extreme case

2 ) N D >> ni :

Because n > nil one then also has N D n >> n:. Under this condition, the
last term in (4.97) may be neglected in comparison with the term n l N D n .
With this, one has the approximate equation

n2 + n1n - n l N D = 0. (4.100)
Its positive solution is given by

(4.101)

If N D << n l , i.e. for relatively light doping, but sufficiently high tempera-
tures, it follows, approximately, that

n =No. (4.102)
All donor atoms are ionized and have donated their additional electrons to
the conduction band. Understandably, for this to occur one needs sufficiently
high temperatures and, maintaining the validity of N D >> nil sufficiently
small donor concentrations. The latter restriction is necessary in order to
4.4. n e e carrier concentrations 487

b
I I I I
tion of temperature for N D =
loi5 ~ r n - (After
~ . Smith, 1979.)
"I
- -- I INTRINSIC RANGE
SLOPE = Eg
7 1
-5
E
Id6
!I
:I
*
t - I SATURATION RANGE

I
$ E
c
; lo"=
I
\
f \nl

1 1
10'3- 1 1 I I 1 I I 1 I I

insure that all donor electrons can be hosted by the conduction band. One
terms this case the exhaustion of doping centers or the saturation of carrier
concentration.
If, conversely, N o >> n1 holds, which means very high donor concentra-
tions and low temperatures, equation (4.101) approximately yields

n= JnTNo. (4.103)
Because of the assumption nl << N o , this concentration is substantially
smaller than N o , i.e., only a small fraction of the donor atoms have been
ionized and have transferred their electrons to the conduction band. One
terms this a reserve of doping atoms or a freez-out of carriers at the im-
purities. Even in this case n can be substantially larger than the intrinsic
concentration ni, provided n1 is larger than or at least of the same order
of magnitude as ni. This occurs for shallow levels with nl >> ni, and is
also still valid for levels E D = ( 1 / 2 ) E g at the middle of the gap for which
nl = ( 1 / 2 ) n i follows. The property of impurity atoms to be effective electron
donors for the conduction band is not restricted, therefore, to the case where
the impurity levels lie just beneath the conduction band edge. Even from
the middle of the forbidden zone they can increase n far above the intrinsic
concentration. This can change only for yet deeper levels with n1 << ni.
The largest increase, however, is caused by shallow levels. This explains
why impurity atoms forming shallow levels are particularly important for
semiconductor doping. Figure 4.5 shows the temperature dependence of the
free carrier concentration in such a case. The three qualitatively different
488 Chapter 4. Electron system in thermodynamic equilibrium

Figure 4.6: Temperature de- 0


pendence of the Fermi level
for n-type Ge assuming typ-
1 EC

ical values of Eg (10 m e v ) ED


and N D c ~ T - ~ )(After
.
Bonch-Bruevich and Kalash-
nikov, 198L)

0 20 40 60 80
T/K -
regions - freez-out of carriers for low temperatures, saturation for medium,
and intrinsic behavior for high temperatures - may be clearly recognized in
this figure.
The minority carrier concentration p in case 2) is always substantially smaller
than p i of the intrinsic case.
We have yet to determine the position of the Fermi level in the case with
considerable n-type doping where N D >> ni holds. Employing the relation

(4.104)

we obtain, for N D << n1, i.e. for high temperatures and low doping, the
result
(4.105)

and, for N D >> nl, i.e. for low temperatures and high doping, we find that

(4.106)

The dependence of the Fermi level on temperature is shown in Figure 4.6 for
n-type germanium with typical values of N D and EB. At T = 0 K the Fermi
level lies exactly in the middle between the donor level and the conduction
band edge, the donors being only partially ionized. With increasing temper-
ature the degree of ionization increases. In order for this to be possible, the
Fermi level must shift to lower energies. This may again be understood as a
repulsion of the Fermi level by the group of states with higher density - the
effective conduction band density of states is larger than the donor density
of states, which equals the donor concentration.

p-type semiconductors

All the above considerations concerning carrier concentrations for n-type


semiconductors can be transferred without difficulty to semiconductors which
4.4. free carrier concentrations 489

contain acceptors instead of donors, i.e. for which N A # 0 and N D = 0.


They are called p-type semiconductors. The holes are the majority carriers
in this case, and the electrons are the minority carriers. We omit details of
derivation here and only display the results. In place of equations (4.92),
(4.93), and (4.94), we have the relations

(4.107)

(4.108)

where E B denotes the acceptor binding energy, and p i is given by

p l = 2 N , e - E /kT. (4.109)
With p instead of n as independent variable, the neutrality condition (4.71)
takes the form

(4.110)

From this, we obtain the equation for p as

P3 + p i p 2 - (ni2 + p i N A ) p - n i2p 1 = 0. (4.111)


For N A << ni the solution is

P = ni, (4.112)
which results in n = ni. For N A >> ni we obtain

P=NA (4.113)
if NA << p l , and
P = & K (4.114)
if N A >> p l . In both cases the minority carrier concentration n is much
less than p . For the Fermi level position, we have results analogous to those
obtained for an n-type semiconductor.

4.4.5 Compensation of donors and acceptors


Above, we considered real semiconductors which contained only donor or
only acceptor atoms, but not both kinds of atoms simultaneously. Now,
we consider the presence of both types of impurities in comparably large
concentrations. Such a state would hardly be created intentionally, but it
may occur unintentionally in the fabrication of a semiconductor material,
490 Chapter 4 . Electron system in thermodynamic equilibrium

where unwanted chemical pollution cannot be avoided. Such pollution could


involve, for example, some boron atoms in a silicon crystal which one would
like to n-dope by means of phosphorus. Also, technological aspects of the
fabrication of semiconductor devices can play a role, for instance, a p-doping
of partial areas of a previously n-doped silicon wafer causes the simultaneous
occurrence of donors and acceptors.
The question arises as to how carrier concentrations adjust under these
circumstances in the semiconductor. Do the respective electron and hole
concentrations simply add without interfering with each other, or do they
strengthen or compensate among themselves completely or partially? A gen-
eral answer to this question can be readily obtained from the mass action
law (4.85). According to this law, the addition of acceptors to a mate-
rial which previously contained only donors, cannot lead to an increase of
the hole concentration without simultaneously lowering the concentration of
electrons. The acceptors therefore compensate the effect of the donors to
a certain degree. The physical reason for this is illustrated in Figure 4.7.
The introduction of acceptor levels slightly above the valence band edge in
a material which already contains donor levels slightly below the conduction
band edge, changes the ground state of the crystal. Instead of transfer-
ring into the conduction band accompanied by a small energy increase, the
donor electrons will transfer into the empty acceptor levels with a substan-
tial lowering of the total energy of the crystal, and the acceptors on their
part preferentially trap these electrons rather than those from the valence
band, since the latter must first be excited. Analogously, holes will move
from acceptor levels to the donor levels instead of being excited into the
valence band. The capture of the free carriers can be complete, whereupon
one speaks of complete compensation If the concentration of acceptors is
too small to host all donor electrons, one has a partially compensated n-type
semiconductor (p < n < N D ) , and if the concentration of donors is too
small to capture all acceptor holes, one has a partially compensated p-type
semiconductor (n < p < N A ) .
In regard to occupancy statistics, the source of electrons (valence band or
donor levels) which occupy the acceptor levels is completely irrelevant. What
does matter is only that levels which can be occupied by electrons exist below
the donor levels. Instead of shallow acceptor levels these can also be deep
levels which do not at all act as acceptors because of their overly large energy
separation from the valence band edge. If they can bind electrons, they
can still be important for the compensation of donors - acting as trapping
centers for electrons or simply as electron traps. Analogous statements hold
for the compensation of acceptors by deep levels which can bind holes, one
has hole traps. As we have seen in section 3.5, deep levels are caused both
by structural defects as well as by chemical impurities. Since semiconductor
materials always contain such perturbations - in larger numbers for lower
4.4. fiee carrier concentrations 49 1

Figure 4 . 7 Semiconductors with a) a donor level, b) an acceptor level, and c)


with both kinds of levels. In parts a) and b) the upper drawings correspond to
the ground state and the lower to the excited state. In part c) the lower drawing
corresponds to the ground state, the acceptor level has captured the donor electron.

cost fabrication - the complete or partial compensation represents, so to say,


the usual state of a semiconductor material. To overcome this requires strong
efforts in material cleaning and in crystal growth. But, even then, certain
semiconductor materials still resist intentional doping. Known examples
occur among the wide-gap 11-VI compound semiconductors, such as ZnS
and ZnSe, which can be n-doped with relative ease, but p-doped only with
difficulty. This problem was solved only recently, when the application of
modern epitaxial growth techniques made it possible to also fabricate p-
doped 11-VI-semiconductorswith wide energy gaps.
Quantitatively, the free carrier concentrations n , p in the case of both n-
and p-type doping may be obtained from the same relations which were used
in the case of pure n- or p t y p e doping. The Fermi level again follows from
the neutrality condition (4.71), which here takes the form

n: n1
n---- N A = 0, (4.115)
n n+n1 N D + nf +pin

where n and p are to be taken from equations (4.79) and (4.80). The relation
(4.115) can be transformed into a fourth-order equation for n or p. Since
this is not solvable in compact form, we consider certain limiting cases which
allow for simple solutions and provide some physical insight. Under the
assumption n << n1, p << pl we find, approximately,

n2 - ni2 - n ( N o + N A ) = 0. (4.116)
This equation corresponds to the limiting case of completely ionized donors
and acceptors. Its solution is given by
492 Chapter 4. Electron system in thermodynamic equilibrium

Figure 4.8: Calculated tempera-


ture dependence of the electron con-
centration of partially compensated T
t OI6
E
n-type germanium. ( E B = 10 meV,
mt = 0.25 m ) . An electron concen-
<
C
1015

tration N D - N A = 10l6 is
assumed. The various curves corre- 104
spond to different degrees of com-
pensation N A : 1- 0, 2 - 3-
loL5, 4 - loL6 (After Bonch-
Bruevich and Kalashnikov, 1982.)
0 2 4 6 8 1012
IOOT-/K-L

For N D = N A , it follows that n = ni. The compensation is complete and we


have the intrinsic case. For N D # N A and ( N o - N A ) >>
~ n?, the result is

n?
p = (NA- No), n= if N A > ND. (4.119)
(NA- N D )
Thus a partially compensated n-type semiconductor occurs for N D > N A ,
and a partially compensated p t y p e semiconductor occurs for N A > ND.
In both cases the majority carrier concentration is J N D- NAJ. For not
completely ionized donors or acceptors, this simple relationship is no longer
valid. In Figure 4.8, partial compensation as a function of temperature is
illustrated using the example of n-doped germanium.

4.4.6 More complex cases

Several different single donors and acceptors


In the analysis above, we considered at most two kinds of simply ionizable
doping centers simultaneously. But, in fact, more can occur, of course. It is
not difficult to expand the theory to this more general case. Each additional
type of a simply ionizable donor (described by an index j) introduces an
4.4. f i e e carrier concentrations 493

+
additional term N D j / ( n n j ) into the neutrality condition, and each addi-
tional type of such an acceptor (described by an index k ) gives rise to an
+
additional term NAkn/(n: n p k ) . The neutrality condition for this more
general case therefore takes the form

n2
n - 2 - z ~ D j - nj + ) : N A ~ zpkn = 0. (4.120)
n j n+nj ni + P k n

As before, it determines the position of the Fermi level uniquely.


For the doping centers themselves, excited bound states were excluded up
to now, and we admitted only two charge states, the ground state and the
simply ionized state. If we give up these assumptions, then the neutrality
condition (4.71) changes. This means that the position of the Fermi level
then depends on N D , N A , N,, N u , and T differently than in the case of centers
having excited states. The particulars of this dependence will be analyzed
below. In this, we may use the fact that the relations (4.79) and (4.80)
between the carrier concentrations n , p and the Fermi level remain the same
as for the case of simple impurities.
Excited bound st ates
We consider a donor which, as before, is only simply ionizable, but has
excited levels ED,, n = 2 , 3 , . . . in addition to the ground state level E D
ED^. In section 4.2, the average number < CdmsN D > of~ electrons
~ at
~
such an impurity was calculated. Using equation (4.29) we find that

(4.121)

with y* given by equation (4.28).


Formally, the occurrence of excited bound states modifies the neutrality
condition (4.97) in that nl is replaced by

61 = y*2n1. (4.122)
Since y* < f, and therefore 6 1 < n l , the concentration n of free electrons
is decreased in general. This is readily understandable since the excited
elect,ron states of the donors result in more electrons being bound.
For acceptors with excited bound states one gets analogous results. The
concentration p l is replaced by

$1 = y*-12-lp1. (4.23)
in the neutrality condition (4.111). Now, j1 > p l holds, which means that
the number of free holes is increased. This is immediately clear because
494 Chapter 4. Electron system in thermodynamic equilibrium

the acceptors bind fewer holes if there are excited electron states at the
acceptors.

Multiply ionizable centers

n-type semiconductors

We consider double donors, such as substitutional sulfur atoms in silicon.


+
The average electron number < fi 1 ND-; > at such centers has been
0,
calculated in section 4.2. The result is shown in equation (4.21). To evaluate
the neutrality condition

n- p -N D ~= 0, (4.124)
one needs the hole concentration Ngf localized at the donors. It is given by
the relation

+
Substituting < fi I fiD-l > as given by equation (4.21), replacing the
DI
chemical potential p by E p l Z E bby Eg - E h , and E g by Eg - E i l where
E L and E i l are, respectively, the ionization energies of one electron at the
1- or 2-fold occupied center, it follows that

As before, in the case of simply ionizable donors and non-degenerate carrier


gases, one can eliminate the Fermi energy from this expression by introducing
the electron concentration n given by (4.79). Instead of the one concentration
constant n1, two now occur, namely

Because E h > E i l , we always have n; < n;. In terms of these constants,


equation (4.126) may be written in the form

(4.128)

With this result (and using n!/n for p ) , the neutrality condition (4.124)
takes the form
4.4. Free carrier concentrations 495

(4.129)

It may be transformed in a fourth-order equation for the determination of


n. Because the difficulty of solving this in closed form, we again consider
only limiting cases. Generally, the doping concentrations ND >> ni should
be sufficiently large that the intrinsic term in (4.129) can be neglected. If
one additionally assumes that

n << n;, and with this, also n << na,

which means that the doping should not be too heavy, nor the temperature
too low, then (4.129) approximately yields

n =~ N D . (4.130)
In this case, all donors are 2-fold ionized and one has the case of exhausted
doping atoms. To fulfill the condition n << n7, it is necessary that N D <<
7x7. The exhaustion of a doubly ionizable donor is thus characterized by
an inequality similar to that which formerly was found for simply ionizable
donors .
In the limiting case

n >> n;,n$

which means sufficiently strong doping and not-too-high temperatures, equa-


tion (4.129) approximately results in

(4.131)

Here, we have set n* = na2/n7 for brevity. The positive solution of this
equation is given by

n = n* Jq- I] (4.132)

With the assumption N D << n*, it yields n = N o , i.e. each donor atom is
simply ionized. One has exhaustion of the doping centers with respect to
1-fold ionization, and reserve with respect to 2-fold ionization. This result
can be understood in the following term: In order to fulfill the conditions
ND << n* and ND >> n7 simultaneously, n* >> ni and with this also nz >> nT
must hold. This means that n* << n$ and ND << na, so that n? << N D << na
496 Chapter 4. Electron system in thermodynamic equilibrium

follows. The condition for donor exhaustion is therefore fulfilled for the
2-fold occupied center, but not for the 1-fold occupied center.
A different situation occurs in the case ND >> n*, in which one finds
from (4.132), approximately,

n= Jm. (4.133)
Because nT < n4, the condition ND >> n* also insures that ND >> na. This
means that the reserve case also occurs with respect to the 2-fold occupied
center. In comparison with a one-particle center having an ionization energy
E B equal to the ionization energy EB of the simply occupied two-particle
center, the carrier concentration due to the two-particle center is enlarged by
the factor 2(n$/n;). The 2 is a consequence of the doubling of the electron
number at the non-ionized donors. The factor ( n ; / n ; ) , which exceeds 1,
reflects the fact that the ionization of the 2-fold occupied center has a larger
probability than that of the simply occupied center.

p-type semiconductors

We consider double acceptors such as substitutional zinc atoms in silicon.


The neutrality condition reads, in this case,

p - - n5
-2N~
(4.134)
P P2 + 2PIP + P2
with

(4.135)
Here EL and Eil refer to electrons (rather than holes) and denote, respec-
tively, the ionization energies of the acceptor with, respectively, 2 holes (no
electron) or 1 hole (1 electron) at the center. Because Eil > EL,one always
has pT > pz. The neutrality condition will again be solved in limiting cases.
In these cases the acceptor concentration NA will always be assumed to be
large compared to the intrinsic concentration p i = ni of holes, so that the
intrinsic term in (4.134) can be neglected. We first assume that

p << p ; , and with this, also p << p ;


holds, which means not-too-strong doping and low temperatures. The neu-
trality condition then approximately yields the relation

P 2NA, (4.136)
corresponding to all acceptors being 2-fold ionized, the case of exhaustion.
In the opposite case
4.4. Free carrier concentrations 497

Figure 4.9: Average electron num-


ber q at an Au-atom in Ge as a func-
tion of the position of the Fermi level
(After Bonch-Bruevich and Kalash-
nikov, 1982.)

we find from (4.134) that

(4.137)

P = $%lP;, (4.138)

follows, corresponding to the reserve case. For p << p ; , equation (4.137)


approximately yields

p = NA. (4.139)
All acceptors are simply ionized, and there is exhaustion of the acceptors
with respect to 1-fold ionization, and reserve with respect to 2-fold ionization.

Amphoteric doping centers

Finally, we discuss a special feature of multiply ionizable doping centers. As


mentioned above in section 3.5, there are amphoteric deep centers simulta-
neously acting as donor and acceptor. In silicon, for example, a neutral Au
atom gives rise to a donor level A u ( O / f )in the lower half of the energy gap,
and an acceptor level Au(-/O) in the upper half. In germanium, there is a
donor level Au(+/2+) of the 1-fold positively charged Au(f1) center which
lies just above the valence band edge (see Figure 4.9). The acceptor levels
Au(2 - /--) and A u ( 3 - / 2 - ) of, respectively, the 1- and 2-fold negatively
charged Au(1-) and Au(2-) centers are found in the upper part of the gap.
Whether Au is a donor or an acceptor thus depends on the charge state or,
in other terms, on the occupancy of the Au center by electrons and, with
498 Chapter 4 . Electron system in thermodynamic equilibrium

this, on the position of the Fermi level (see Figure 4.9). The latter is co-
determined by the concentration of Au atoms, and it also depends, however,
on the concentrations of other donors and acceptors if there are such. In
p-type doped germanium, Au acts as donor because, in this case, the Fermi
level lies just above the valence band edge so that the charge state Au(l+)
occurs. In n-doped germanium, Au forms a double acceptor due to the Fermi
level position lying slightly below the conduction band edge, which results
in the charge states Au(1-) and Au(2-).
499

Chapter 5

Non-equilibrium processes
in semiconductors

In the preceding chapter we considered semiconductors which were spatially


homogeneous in a macroscopic sense and were not subject to external per-
turbations or fields. The thermodynamic equilibrium state of electrons in
such semiconductors is characterized in terms of thermodynamic variables,
in particular a temperature and a chemical potential, which do not depend
on space and time coordinates. Semiconductors differ from other materi-
als inasmuch as they can easily be driven out of equilibrium. The electron
temperature and chemical potential may then differ relatively strongly from
their equilibrium values, provided such state variables do exist at all. If this
is the rase, they may depend on space and time coordinates. In addition,
there may be relatively large electric fields and currents. Relatively means
with respect to other solid state materials. In metals, by way of comparison,
macroscopic electric fields practically cannot exist, because they are almost
completely screened out over distances of about 1 A by the huge number
of mobile electrons available in these materials. The same may be said for
spatial variations of the chemical potential in metals, which also practically
cannot occur because of the large electron concentration.
Semiconductors have the unique property that external manipulations
can easily bring them into states which deviate strongly from thermody-
namic equilibrium. It is on this unique property that essentially all ap-
plications of semiconductors in electronic and optoelectronic devices ulti-
mately rest. The semiconductor photoresistor, for example, works because
non-equilibrium electrons and holes are generated by irradiation with light,
greatly changing the resistance of the semiconductor sample. In a field effect
transistor, the resistivity of a particular semiconductor region is changed by
means of an external voltage, and in a light emitting semiconductor diode,
non-equilibrium minority carriers, created at a forward biased pn-junction,
500 Chapter 5. Non-equilibrium processes in semiconductors

recombine with the equilibrium majority carriers accompanied by the emis-


sion of light. While the latter two examples refer to semiconductor junctions,
the following considerations will initially deal with bulk semiconductors, i.e.
semiconductors with spatially constant doping and chemical composition.

5.1 Fundamentals of the statistical description of


non-equilibrium processes
Thermodynamic equilibrium is the only macroscopic state of the electron
system of a bulk semiconductor which does not change in time if one leaves
the semiconductor alone. In all other macroscopic states, i.e., all non-
equilibrium states, the system undergoes temporal changes. This means that
time-dependent macroscopic processes occur if the semiconductor electron
system is out of equilibrium. One terms them non-equilibrium processes,
and it is these which we will discuss in the present chapter.
Within the framework of a semiclassical microscopic statistical theory,
non-equilibrium states of an electron system are described by occupation
probabilities f of one-electron states i which depend not only on the energies
Ei of these states, as in thermodynamic equilibrium, but on the quantum
numbers themselves. In general, they are also functions f ( i , x,t ) of space and
time coordinates x, t . The transition from equilibrium to non-equilibrium
states is thus described by the replacement

Since, ultimately, the temperature T and chemical potential p are defined as


parameters in the equilibrium distribution function f(Ei), these quantities
are not, in general, meaningful for macroscopic non-equilibrium states. Ac-
cordingly, there is also no corresponding non-equilibrium distribution func-
tion f ( i ,x,t ) which would depend on them. Moreover, there is no handy
general expression for f ( i , x,t ) at all which would depend on macroscopic
parameters to be adjusted to the particular non-equilibrium state, in con-
trast to the case of equilibrium. Instead of analytical expressions one may
formulate equations from which the distribution function f ( i , x,t ) can be
determined. Of such equations, the Boltzmann equation is the simplest for
the electron system of a crystal.

Boltzmann equation
We will write this equation down explicitly for the extended one-electron
excitations of the system, i.e. the Bloch states uk. The distribution function
f(uk, x,t ) in this case is the probability that, at a position x and at a time
t , the Bloch state having quasi-wavevector k and band-index u is occupied
5.1. Fundamentals of the statistical description of non-equilibrium processes 501

by an electron. Since, in reality, the Bloch states are infinitely extended, the
specification at a position x must be understood in the sense of a region
centered at x, whose spatial extension A is large compared to the electron
wavelength X = 2 n / J k J .For the concept of a spacedependent distribution
function f(vk, x, t ) of Bloch states to make sense, f(vk, x, t ) must change
very slowly over a length of the order of magnitude A. A similar condition
holds for the variation of the distribution function with time.
In the absence of external perturbations of the type which drive a current
or energy transfer, the system of Bloch electrons is in thermodynamic equi-
librium, such that it may be described by the Fermi distribution function
(4.13) with E = E,(k). To drive the system out of equilibrium, appropriate
external perturbations must be applied, such as particle density gradients or
an external electric field E(x, t ) , as we will assume here. In addition to ex-
ternal perturbations, internal perturbations also act on the Bloch electrons,
such as scattering by phonons and impurities. The term collisions is used
in this context which means that the scattering events take place in a very
short time, practically instantaneous. Both types of perturbations, external
ones and collisions, result in temporal changes of the distribution function,
denoted as (df(vk, x, t ) / d t ) e z t and (df(vk, x, t ) / d t ) , d l , respectively. Accord-
ing to a general theorem of statistical mechanics, the Liouville theorem, the
total change o f f , taken at a position x(t),iik(t) of phase space which moves
together with the particles, must vanish. Thus,

Performing the derivative, the first term of this equation may be written as

Here, (dx(t)/dt)may be replaced by (l/h)V&,(k) according to equation


(2.193), and ( d k ( t ) / d t )by -eE/h. according to equation (3.310). Combining
equations (5.2) and (5.3) yields the Boltzmann equation

[i+ iVkE,(k) . V, - eE(x,t ) .


Cdl

The collision term is in general an integral over products of two distribu-


tion functions, one of which, f(vk, x, t ) , involves the same Bloch state vk
occurring on the left hand side, but the other, vk, involves any Bloch state
which can participate in a collision. The characteristic difficulties of the
502 Chapter 5. Non-equilibrium processes in semiconductors

Boltzmann equation are mainly due to its collision term, a statement which
applies both to the derivation as well as to the solution of this equation.
Lets suppose that the Boltzmann equation has been solved. Then, as in the
case of equilibrium, one may use the known distribution function f(vk, x,t )
to determine statistical average values < A > of any one-particle quantity
A of the many-electron system. Such quantities represent sums over cor-
responding individual electron quantities a. Like the distribution function
itself, the average values are functions < A > (x,t ) of space and time coor-
dinates, such that

< A > (x,t ) = C f ( v k , X , t)(vk)alvk). (5.5)


uk

In addition to the physical quantities which are of interest in both equi-


librium and non-equilibrium, such as the electron concentration and the
energy density, non-equilibrium also involves other quantities to be deter-
mined, in particular, currents related to equilibrium quantities such as, for
example, particle or energy currents. The latter differ from zeroonly in
non-equilibrium, and thus are of no interest in equilibrium. Moreover, those
quantities whose average values do not vanish in equilibrium must also be re-
considered, because under non-equilibrium conditions they exhibit temporal
and spatial variations which are not present in equilibrium. As an example,
one may take the total number of free electrons in a spatially homogeneous
semiconductor sample. Under non-equilibrium conditions this number may
have larger or smaller values than in equilibrium. If one withdraws external
influences, equilibrium and its parameter values will be restored by non-
equilibrium processes.
The two classes of phenomena, the appearance of currents, on the one
hand, and, of time and space changes of quantities which are non-zero but
constant in equilibrium, on the other hand, constitute the totality of non-
equilibrium processes. For semiconductors, both types of processes are im-
portant. Their theoretical description can be based on the Boltzmann equa-
tion - one solves this equation and then employs the resulting distribution
function f(vk, x,t ) to calculate the average values of the various quantities
of interest. Alternatively it can be based on empirical equations for the av-
erage currents and the temporal or spatial change rates. We refer to these
as phenomenological equations. Ohms law for electrical conduction and the
first and second Ficks laws for diffusion are examples.

Phenomenological equations and their derivation from the


Boltzmann equation

Insofar as information about the state of the system is concerned, the Boltz-
mann equation provides much more detail than the phenomenological equa-
5.1. Fundamentals of the statistical description of non-equilibrium processes 503

tions. The latter can be derived from the Boltzmann equation with addi-
tional assumptions, but the reverse, i.e. the derivation of the Boltzmann
equation from the phenomenological equations, is not possible. The transi-
tion from the Boltzmann equation to phenomenological equations represents,
essentially, a coarse graining of the length and time scales on which the non-
equilibrium processes proceed. If processes must be analyzed on the space
scale of the average mean free path length and the time scale of the mean
free flight time, then Boltzmanns equation is required. If a macroscopic
average description suffices, one can use the phenomenological equations. In
semiconductor physics, the latter case often applies. Accordingly, we will r e
strict our considerations to the phenomenological equations in our treatment
of non-equilibrium processes.
These equations are valid only within prescribed limits. One can con-
sider these limits, like the equations themselves, to be empirically given.
However, deeper insight can be gained by using Boltzmanns equation. The
question to be answered in this context is: Under what assumptions can
the phenomenological equations be derived from the Boltzmann equation?
The answer depends on the phenomenon under consideration. Equations as,
for example, that for the temporal change rate of a given physical quantity,
may be obtained by multiplying the Boltzmann equation by the value of this
quantity in a Bloch state vk and subsequently summing over all states. The
simplest equation of this type is that for the temporal change rate of the free
carrier concentration of a single band, which later will explicitly be written
down. To obtain this equation from Boltzmanns equation in the manner
indicated above, simplifying assumptions are necessary concerning the quan-
tum mechanical transition probabilities between bands and localized states
occurring in the collision term: These probabilities must be approximated
by their values in thermodynamic equilibrium.
In the derivation of phenomenological equations for particle currents,
to consider another example, one assumes that the non-equilibrium state
deviates only slightly from a spatially and temporally local equilibrium state.
In the latter state, the distribution function depends on the Bloch states only
through their energies, and the form of the dependence is that of the Fermi
distribution function f ( E ) , just like in global equilibrium. However, the
temperature T and chemical potential p occurring in f(E), are understood
as functions T(x,t ) and p(x, t) of space and time coordinates. Thus we have,

This particular non-equilibrium state may only be assumed if the tempera-


ture and chemical potential are still well defined quantities, albeit in a local
sense, not globally. For this assumption to be valid, adjacent parts of the
504 Chapter 5. Non-equilibrium processes in semiconductors

system with measurably different values of temperature and chemical poten-


tial are effectively decoupled from each other spatially and temporally. From
the microscopic theory of non-equilibrium processes it is known that spatial
decoupling (in the statistical sense of this term) occurs over a distance of
the order of magnitude of the mean free path length l f , and temporal decou-
pling occurs over a time interval of the order of magnitude of the mean free
flight time t f . The characteristic lengths of the spatial inhomogeneities of
T(x, t ) and p(x, t ) may be expressed by the ratio TIIVTI and p/IVpL(, and
the characteristic time of the temporal inhomogeneities of these quantities
is given by T/laT/atl and p/lOp/OtL(. In order for local equilibrium to be
approximately realized, the conditions

If << TIIVTI, P I P P I ?
and
tf << TllaTlatl, P l l a P l a t l (5.7)
must be satisfied. Typical values for the mean free path length of electrons
and holes in semiconductors are of the order of magnitude 100 A. This means
that local equilibrium is a reasonable zero order description if important
spatial inhomogeneities occur on or above a length scale of 1000 A = 0.1 p m .
The mean free flight time has typical values close to s. Temporal
changes with frequencies of 10l2 s- can therefore still be accepted within
the local equilibrium framework, although no light frequencies are allowed
which lie in the region of 1014 s-. The regime of validity of the local
equilibrium assumption is nevertheless still relatively broad.
If the local equilibrium distribution were to describe the non-equilibrium
state exactly, no currents would flow at all since the corresponding average
values would be zero. In the presence of currents, the local equilibrium
distribution function can, therefore, only be a zero-th approximation. It
must be corrected by a perturbation term fl(vk, x, t ) , such that one has

For an electric current driven by an electric field E ( x , t ) , Ohms law rep-


resents the relevant phenomenological equation. To obtain this law, the
correction term is written as a linear relation

with respect to the electric field E(x, t ) . The vector g(vk, x, t ) is understood
to be field-independent. It may be calculated from Boltzmanns equation in
linear approximation with respect to E. For this approximation to be justi-
fied, the field must be sufficiently weak. This represents a second condition
for the validity of Ohms law.
5.2. Systematics of non-equilibrium processes in semiconductors 505

The above derivation of this law yields not only a phenomenological equation,
but also an expression for the electrical conductivity involved in it. It relates
this coefficient with characteristic microscopic quantities of the system such
as, for example, its band structure, as well as with external state parameters
like temperature and chemical potential. This is to say that derivation from
the Boltzmann equation yields the phenomenological equation, Ohmss law,
and, simultaneously, a microscopic theory for the conductivity, or, if other
phenomenological equations are considered, for the other respective kinetic
coefficients. If the phenomenological equations are taken as empiricaI laws,
these coefficients are material parameters whose magnitudes and tempera-
ture dependencies can only be determined from experimental investigations.
However, by means of Boltzmanns equation the kinetic coefficients can be
predicted. Without the use of such an equation, this opportunity is lost. Be-
side the somewhat lower precision of the phenomenological equations, due
to simplifications and approximations, the principal loss lies in the inability
to predict the kinetic coefficients, given only phenomenological equations in-
stead of Boltzmanns equation. For the purposes of this book, where we do
not intend to develop a detailed understanding of the magnitudes and state
dependencies of material parameters, we will accept this loss and use the
phenomenological equations to gain insight into non-equilibrium dynamics
quickly.
In the next section we present, first of all, a systematic survey of the non-
equilibrium processes which are essential in semiconductors. More detailed
treatment of each process is provided in following sections.

5.2 Systematics of non-equilibrium processes in


semiconductors
The following survey of non-equilibrium processes is based on a gradual
introduction of temporal and spatial inhomogeneities into the semiconductor.
In the first step we admit only a temporal inhomogeneity, while spatial
homogeneity will be maintained. The non-equilibrium processes which then
occur are commonly called relaxation processes.

5.2.1 Temporal inhomogeneity and spatial homogeneity


In this case, all state quantities, particularly the charge carrier concentra-
tions n and p , depend only on time, and their values are different from those
in thermodynamic equilibrium. The latter will be denoted by no and po
henceforth, so that

(5.10)
506 Chapter 5. Non-equilibrium processes in semiconductors

with
(5.11)

Temporal changes of n and p occur because electrons in the conduction


band and holes in the valence band are generated or annihilated. If the
corresponding generation or annihilation rates are denoted by, respectively,
(an/at)gmer and (an/at)annihal as well as (aP/ at)gmer and (aP/at)annihal,
we have

(%) = -

(5.12)

With the right hand sides known, these equations determine the relaxation
of non-equilibrium charge carrier concentrations.

5.2.2 Spatial inhomogeneity and temporal homogeneity


Spatial inhomogeneity may be introduced in various ways. Firstly, we con-
sider the case in which a homogeneous electric field exists while the charge
carrier concentrations n and p do not depend on x:

1) n = no, p = p o , E = c o n s t . # 0.
The potential energy of an electron is eE . x in this case, i.e. the system
is in fact spatially inhomogeneous despite its homogeneous charge carrier
concentration. An electric current will flow in this case. This phenomenon
is commonly referred to as drift, and the current itself is called drift current.
The pertinent current density is spatially homogeneous and decomposes into
a current density for electrons, j,, and another one for holes, j,. With nn and
a, as conductivities for electrons and holes, respectively, Ohms law yields

j, = onE, j, = apE. (5.13)


Secondly, we also admit spatial inhomogeneity of the charge carrier concen-
trations n , p . Then the electric field also becomes inhomogeneous. Accord-
ingly, we consider the case

2) n = n(x), p = p(x), E = E(x) # 0.


Under these conditions one has, in addition to drift, also a diffusion current
of charge carriers which is superimposed on the drift current. Writing Dn
and D , as diffusion coefficients for, respectively, electrons and holes, the
5.2. Systematics of non-equilibrium processes in semiconductors 507

total electron and hole current densities j, and j, obey the phenomenological
equations

jn(x)= o,E(x) + eD,Vn(x), (5.14)

j,(x) = oPE(x)- eD,Vp(x). (5.15)

Yet another new phenomenon arises in the inhomogeneous case considered


here, namely the formation of space charge, reflected in the occurrence of
a non-vanishing local charge density p(x). To describe this, we recall the
neutrality condition n - - p - N & + N i = 0 derived in Chapter 4. It states that,
in a homogeneous semiconductor, the negative charge density of the freely
mobile electrons and of the spatially fixed ionized acceptors compensates the
positive charge density of the holes and the ionized donors everywhere. In
the spatially inhomogeneous semiconductor considered here the free electrons
and holes are redistributed while the ionized doping atoms are kept fixed, so
that local charge neutrality is perturbed. The local net charge density p(x)
is given by the expression

P(X) = --e [.(XI +


- P(X) - Nof(x) NJX)] , (5.16)

where we have allowed for the possibility that a spatially inhomogeneous


distribution of ionized donors and acceptors also exists, either due to inho-
mogeneous doping or to inhomogeneous ionization of the impurity atoms.
Charge neutrality holds only for the semiconductor as whole,

s, d3xp(x) = 0. (5.17)

The electric field E(x) due to the charge distribution p(x) follows from the
pertinent potential p(x) by means of the relation

E(x) = -Vp(x), (5.18)


and the potential p(x) itself is governed by Poissons equation

1
V2p(x) = --p(x) (5.19)
=0

with E as the static dielectric constant of the semiconductor material and 0


the vacuum dielectric constant.
Finally, we consider in our survey of non-equilibrium processes the most
general case in which both space and time inhomogeneities exist.
508 Chapter 5. Non-equilibrium processes in semiconductors

5.2.3 Space and time inhomogeneities


In this situation we have

n = n ( x ,t ) , p = p ( x ,t ) , E = E ( x , t ) # 0.

Here, the generation and annihilation of charge carriers is also possible be-
cause of the fact that there are sources and sinks of the current densities j n
and j,. This follows from the continuity equations for j n and j , , where the
rates of change of the charge carrier concentrations due to relaxation enter
as source terms. Thus. we have

The current densities continue to obey the phenomenological equations] even


with temporal variation in the present case, so that

j p ( x ,t ) = u p E ( x ,t ) - ~ D , ~ P (t )x , (5.23)

The Poisson equation for the potential p ( x , t ) reads

Summing up, we have the following non-equilibrium processes in semicon-


ductors:

- Generation and annihilation of free charge carriers (relaxation)


- Drift
- Diffusion
- Formation of space charge and of the concomitant inhomogeneous elec-
tric field and potential.

Below we treat each of these non-equilibrium processes in greater detail.


5.3. Generation and annihilation of free charge carriers 509

5.3 Generation and annihilation of free charge


carriers

Since the total number of electrons in a semiconductor is conserved, free


charge carriers can neither be produced without a well defined origin, nor
can they vanish without a well defined destination. The generation of free
charge carriers in a spatially homogeneous semiconductor must therefore be
understood in the sense that electrons or holes, which are not freely mobile
but bound to impurities and defects, are excited into the conduction and
valence bands where they can move freely. The generation of free charge
carriers thus takes place at the expense of localized carriers that are not
free to move. The same is true for the annihilation of free carriers. They
do not simply disappear, but are captured by states localized at impurities
or defects in which they cannot move freely. According to Chapter 4 the
free carriers of a semiconductor in equilibrium are mainly provided by ther-
mally ionized shallow donor or acceptor centers. However, for generation
and annihilation of non-equilibrium carriers these centers play only a mi-
nor role if their ionization is almost complete under equilibrium conditions.
In Chapter 4 we saw that this is in fact often the case, mainly because the
characteristic thermal energy kT is comparable to the binding energies of the
shallow centers. One may expect that deep centers should be more effective
for the generation and annihilation of non-equilibrium free carriers than are
the shallow ones. Under equilibrium conditions, deep centers are generally
in their ground state, in which they are occupied by a number n of electrons
(see section 3.5). By the application of an external perturbation, such as
light, for instance, they can be transferred into excited states, in which they
have exchanged electrons and holes with the valence and conduction bands.
Such an exchange is, however, exactly the generation and annihilation of free
carriers in which we are interested. Deep centers are therefore important for
the generation and annihilation of non-equilibrium carriers.
Moreover, valence band states can play a role similar to that of deep
levels for non-equilibrium electrons to be generated or annihilated, and con-
duction band states can do so for non-equilibrium holes. In considering this,
one must bear in mind the fact that electrons in the valence band and holes
in the conduction band are also immobile charge carriers. The excitation
of electrons from the valence band into the conduction band generates free
carriers in the latter. However, it simultaneously creates holes, i.e., free
charge carriers in the valence band, which stands in remarkable contrast to
the generation of free electrons by exciting deep levels, in which case the r e
maining hole is immobile. An analogous statement holds for the annihilation
of free electrons. When an electron falls back from the conduction band into
the valence band, then a free hole simultaneously vanishes together with the
5 10 Chapter 5. Non-equilibrium processes in semiconductors

free electron, while the transition of a conduction band electron into a deep
level only removes a free electron. A free hole, which could vanish, does not
exist in the latter case. If free electrons and holes are simultaneously gen-
erated or annihilated in equal numbers, we speak of bipolar generation and
annihilation processes. If only one type of free charge carrier is generated or
annihilated, we have unipolar processes. We first treat generation processes,
including uni- and bipolar ones.

5.3.1 Generation processes


To excite free charge carriers from deep levels or from bands, an appropriate
amount of energy must be provided. That can be done by exposing the
semiconductor to light or electron beams, and in many other ways. The
creation of free carriers by applying an external voltage to a semiconduc-
tor junction, so-called injection, is excluded here because we are confining
our attention to spatially homogeneous semiconductors. Injection is, by its
nature, a unipolar process and it will be treated later. Of the remaining
processes, excitation by means of light will be chosen as a representative
example. We refer to it as optical excitation, and other excitation processes
can be described in similar terms.
Unipolar optical excitation processes of electrons and holes are illustrated
in Figure 5.1. An electron from a deep center, which in thermodynamic equi-
librium is in its ground state, is raised into the conduction band by a photon
of proper energy. As a result, a free electron appears in the conduction band.
Alternatively, an electron from the valence band makes a transition into a
deep level with the absorption of a photon. A free hole is generated in the
valence band as a result of this process. We denote the number of photons
absorbed per volume and time unit by, respectively, gn and g p . Then the
rate of unipolar-generated carriers may be expressed as

(g) unip.gen.
=gnI
unip.gen.
= gp. (5.25)

In the case of bipolar generation by means of light (see Figure 5.1) one has
gn = g p = g , so that

(5.26)

To estimate the magnitude of the optical generation rates, we consider the


bipolar case. With I as light intensity and a as the absorption coefficient
associated with the excitation of electrons from the valence band into the
conduction band. we have
9= (E). (5.27)
5.3. Generation and annihilation of free charge carriers 511

a)
I
.
...
.

Figure 5.1: Unipolar generation of electrons (a), holes (b), and bipolar generation
of electron-hole pairs (c).

For light intensity of 1 Wcrn-, photon energy hw = 1 e V , and absorption


coefficient lo3 ern-, which is not too large for the absorption edge of a
semiconductor, we obtain g = 6 x loz1 ~ r n - ~ s - . This indicates that op-
tical excitation represents a very effective generation process. The size of
g does not tell, however, how large the steady value of the free charge car-
rier concentration will actually be. This value also depends on the speed of
annihilation processes.

5.3.2 Unipolar annihilation of free charge carriers: capture


at deep centers
As outlined above, free charge carriers are annihilated when they make a
transition into deep levels. One refers to this process as the capture of free
charge carriers by deep centers (see Figure 5.2). We first consider the capture
of electrons, and assume that the deep center D can bind one and only one
more electron than it has in its ground state. If the center is neutral in its
ground state, this means that the D(-1/0) donor level lies in the gap, while
the 0 ( - 2 / - 1) level does not appear there. For brevity, the ground state of
the deep center will be referred to as unoccupied, and the state with one
electron more at the deep center as occupied. The ionization energy of the
occupied center is denoted by Et.

Electron capture
After an electron of the conduction band is captured by a deep center, there is
a finite probability that it may be re-emitted back into this band. The anni-
hilation rate -(an/dt),,,ihii is therefore the net capture rate (dn/dt),pture,
defined as the negative of the difference between the gross capture rate C,
and the emission rate En, i.e.
5 12 Chapter 5. Non-equilibrium processes in semiconductors

Figure 5.2: Capture (on the


left) and emission (on the right)
of electrons (above) and holes
(below) at a deep level. .
...
...
..
-- Y EP

(5.28)

The capture rate C , is, on the one hand, a function of the concentration of
free electrons available for capture, i.e. of n. On the other hand, it depends
on the concentration of deep centers available for a capture. If we denote
by N t the total concentration of deep centers, and by f the probability that
they are already occupied by an electron, then the concentration of available
centers is Nt(1 - f ) . Thus,

c, = c,n(l - f)Nt. (5.29)


The proportionality factor c, is called capture coefficient. It has the di-
mension cm3s-' and describes the capture capability of an unoccupied deep
center.
The emission rate En depends on how many deep centers are already
occupied by electrons, i.e. on Nt f . The concentration of existing free elec-
trons has no influence on En, since the number of unoccupied states in the
conduction band which must host the emitted electron is relatively large.
Therefore,

En = en f Nt. (5.30)
The proportionality factor en is called emission coefficient. It measures
the emission capability of an occupied center. For one occupied center the
emission probability per second is just en. The net capture rate follows as

($) capture
= - [cnn(l- f ) N t - e n f ~ t ~ . (5.31)

Consider first the special case of thermodynamic equilibrium. In this state all
macroscopic quantities are independent of time, so that (dn/bt)capture = 0
holds. The concentrations n , p take their equilibrium values no and PO, and
5.3. Generation and annihilation of free charge carriers 513

the same holds for the occupation probability f of the deep level, which is
given by

(5.32)

where -ytn denotes the ratio of the statistical weights of the unoccupied and
occupied center. Equation (5.31), with vanishing left hand side, yields a
relation between en and cn in thermodynamic equilibrium, namely

with
(5.34)

The quantity nt has a structure similar to that of the concentration nl


previously introduced in Chapter 4. It denotes, apart from the factor Ytn,
the concentration of free electrons in the case when the Fermi energy lies
exactly at the deep level. For an n-type semiconductor, EF is actually found
to be much closer to the conduction band, i.e., nt is, in general, substantially
smaller than the equilibrium concentration no.
The relation (5.33) between emission and capture coefficients is a con-
sequence of the equilibrium between the deep level and conduction band.
The term detailed balance is used in this context, since the two levels are in
equilibrium already themselves.
If there are deviations from equilibrium, the net capture rate (5.31) is
not fully determined inasmuch as the occupation probability f of the deep
center has not yet been further specified. In the general case, this probability
will also deviate from the equilibrium value fo. The exact value which it will
have cannot be predicted a priori. One can only write down an equation
for its temporal rate of change. With the assumption that the exchange of
electrons occurs only between the deep level and the conduction band, this
equation reads

(5.35)

If we also assume that beyond a particular point in time t o no further gener-


ation of free charge carriers will take place, but only annihilation processes
will occur through capture , we have

(g) (g) = capture


for t > to. (5.36)

The relations (5.35) and (5.36) together form a system of non-linear differen-
tial equations, which can be solved exactly in closed form. However, we will
514 Chapter 5. Non-equilibrium processes in semiconductors

not exhibit this solution here, restricting our discussion to a linear approxi-
mation. In this, we assume that the functions n and f deviate only weakly
from their equilibrium values. Formally, this means that n = no An and +
+
f = f o A f with lAnl << no and lAf I << 1, whence

(g) capture
= - c n N [-(nt + no)Af + An(1- f o ) ] . (5.37)

Bilinear terms were neglected in this derivation, and equation (5.33) was
used. We consider that the generation of non-equilibrium electrons before
time t o arises from an initial equilibrium distribution with subsequent exci-
tation of electrons from deep levels into the conduction band. The resulting
total concentration of electrons in deep levels and in the conduction band
+
jointly have the equilibrium value no Ntfo at all points of time t before
and after to. Therefore

An + NtAf = 0. (5.38)
With this, the linear expression

(5.39)

follows from (5.37), where for the sake of brevity we have defined
1
7, = (5.40)
cn[nt + no + N t ( 1 - f o ) ]

The time development of An is governed by the equation


An
(5.41)

The solution of this equation reads

where An0 = An(t = t o ) has been used. This result demonstrates that rn
represents the decay time of a non-equilibrium charge carrier distribution.
One also calls rn the lifetime of non-equilibrium electrons with respect to
capture at deep centers. In the case of completely unoccupied deep centers,
i.e. at fo = 0, and for very high concentrations Nt, more exactly for Nt >>
+
nt no, rn is given by
1
I
rno = -. (5.43)
CnNt
5.3. Generation and annihilation of free charge carriers 515

In accordance with its meaning, rn0 is the time interval in which the totality
of Nt unoccupied deep level per cm3 capture one non-equilibrium electron
from the conduction band. One therefore calls r,,o the capture time. The life-
time r,, and the capture time rno differ because of the occupation dependence
of the capture rate and the occurrence of reemission. If N t ( 1 - fo) >> nt+no,
then one has r,, > rno. This result is immediately understandable because
the lifetime r,, contains contributions only from unoccupied deep centers,
whereas all deep centers, including occupied ones, enter in rno.

Hole capture

As in the case of electrons, deep centers can also capture and reemit holes.
The capture of a hole means that a deep center, which in its ground state
binds a certain number of electrons, transfers one of these electrons to the
valence band and annihilates a hole there. The emission of a hole is to
be understood in the sense that a deep level captures an electron from the
valence band (see Figure 5.1). All considerations involving the capture and
emission of electrons, and all relations derived for these processes, can be
immediately transferred to holes. The capture and emission rates C p and
E p are given by the expressions

In this, cp and e p denote the capture and emission coefficients for holes.
They are related to each other by the equation

ep = PtCp 7 (5.45)
where pt is the hole concentration

(5.46)

The meaning of ytp is analogous to that of y h . The hole lifetime rp obeys


the relation
1
rp = (5.47)
Cpbt + P O + Ntfol '
and for the hole capture time ~~0 of a deep center we have

(5.48)
516 Chapter 5. Non-equilibrium processes in semiconductors

Table 5.1: Experimental capture cross sections for particular deep centers in Si
and Ge at T = 300 K in units cm2.(Data compiled from various sources.)

n Si
Center on UP

Au(l+) 35-63
Au(0) 1.7-5
Au(1-) 10-110
140) 0.1 Cu(1-) 0.1-0.18 500
In(l-) 170 Cu(2-) 0.0036-0.3 1-3.6

Bi(0) 0.025 Ni(0) 1-8


Bi(1-) 0.6 Ni(1-) 3-6 1000
Bi(2-) 0.6-1 Ni(2-) 5 - 20 x 1000
Fe(l+) 16 Fe(0) 1-10
WO) 3-7 Fe( 1-) 30
Fe(2-) 100

Zn(0) 10 Zn(1-) < 0.001


Zn(1-) 0.0001-1 300
Ga(0) < 0.01 Ga(0) < 0.01

Experimental determination of capture coefficients

Experimentally, the emission coefficients e , of deep centers can be deter-


mined by the same time-resolved current or capacitance measurements at
pn-junctions that were discussed in section 3.5. The capture coefficients c,
follow from en by means of equation (5.33). The experimental data for the
capture efficiency of deep centers are generally expressed as capture cross-
sections a, and up. These are related to the capture coefficients c, and
cp by a simple relation. With v n and up as average thermal velocities of,
respectively, electrons and holes, we have

(5.49)

Experimentally determined capture cross sections for some deep centers are
listed in Table 5.1.
5.3. Generation and annihilation of free charge carriers 517

Finally, we present a short discussion of the microscopic mechanisms respon-


sible for the capture of free charge carriers at deep centers.

Capture mechanisms

We restrict our considerations to electrons, as the transfer of the discussion


to holes is evident. To understand capture processes, the first question that
has to be answered is how an electron, during its transition from the con-
duction band into the deep level, can transfer away its surplus energy. There
are essentially three possibilities. Firstly, the energy can be emitted as light,
secondly, it can be transferred to the remaining electrons of the semiconduc-
tor, and thirdly, it can be transferred to the lattice. The last case is the most
important and most universal. The transfer of electron energy to the lat-
tice involves the interaction of the electrons with phonons. This interaction
was ignored in Chapter 2 by treating the interacting system of electrons and
atomic cores within the adiabatic approximation. In order to account for the
electron-phonon interaction, one must go beyond this approximation. More-
over, it must be recognized that one single phonon is insufficient to make
possible the transfer of energy during capture, and many phonons must
be emitted simultaneously. This is termed a multi-phonon process. Conse-
quently, a perturbation theoretical treatment of the electron-phonon interac-
tion is not adequate for this purpose, and one must employ non-perturbative
calculations of the capture coefficients due to multi-phonon processes at deep
centers (see Peuker, Enderlein, Schenk, and Gutsche, 1982).

5.3.3 Bipolar annihilation of carriers at deep centers


If an electron of the conduction band is annihilated by a transition into a
hole of the valence band, we have a recombination of a n electron-hole pair.
The physical mechanisms for the transfer-away of the surplus energy, which
in this context is at least as big as the gap energy E,, are largely similar
to those for the capture of carriers at deep centers. Firstly, an electron can
make the transition from the conduction band into a hole in the valence
band by emitting a photon. This process is called an optical band-band re-
combination The term radiative recombination is also commonly used. It is
known that spontaneous optical band-band transitions, i.e. transitions not
stimulated by light, have only a relatively small probability. Often, they can
be completely neglected. Optically stimulated processes, which play a cru-
cial role in semiconductor lasers, are considerably more effective. Secondly,
the surplus energy of a recombining electron-hole pair can be transferred to
another electron in the conduction band, which is thereby lifted to higher
energy in the band. In this process, referred to as A u g e r recombination, three
free charge carriers are involved, two electrons in the conduction band and
5 18 Chapter 5. Non-equilibrium processes in semiconductors

one hole in the valence band. Auger recombination is important mainly for
high charge carrier concentrations, since its probability is determined by the
product of the concentrations of the three carriers involved. Thirdly, the en-
ergy released during recombination can be transferred again to the lattice.
This direct band-band recombination with emission of as many phonons
as needed to match the energy gap, proves to be extremely unlikely. The
reasons for this are, on the one hand, that the number of phonons which
would have to be emitted simultaneously is very large, and secondly that
the electron-phonon interaction for electrons and holes in band states is rel-
atively weak. Recombination processes in which deep levels are involved
as intermediate states for an electron making the transition from the con-
duction band to a hole in the valence band are much more effective. This
mechanism is called Shockley-Read-Hall recombination Auger recombina-
tion and Shockley-Read-Hall recombination together are jointly referred to
as non-radiative recombination processes. Because of its dominating role in
electronic devices, we treat Shockley-Read-Hall recombination below in more
detail.

S hockley-Read-Hall recombinat ion

For a deep center X to be effective in mediating the recombination of an


electron-hole pair, it must be able to capture both an electron and a hole.
This means that both the X(-l/O)-donor level and the X(-l/O)-acceptor
level of the center must lie in the gap. The capture of a hole by a deep level
entails the emission of an electron into the valence band. The simultaneous
capture of an electron and a hole (see Figure 5.3) corresponds, therefore,
to the transition of an electron from the conduction band into a hole in the
valence band, i.e. to the recombination of an electron-hole pair. The occupa-
tion state of the deep center after recombination has taken place is the same
as that before recombination - it is in the ground state, the captured elec-
tron and the captured hole have compensated each other. Nevertheless the
deep center plays a decisive role in the recombination process. It enhances
its probability by many orders of magnitude. The deep center works, so to
say, as a catalyst for recombination. The reason for this follows immediately
from our discussion of the multi-phonon processes above.
A deep center not only captures electrons and holes, but it also emits
them (see Figure 5.2). In order that a deep center actually function as a ten-
ter for recombination, a particular relation between the capture and emission
rates C,, C, and En ,Ep for, respectively, electrons and holes, must hold.
On the one hand, an electron must be captured from the conduction band
more quickly than an electron from the valence band. If that were to be
reversed, the electron from the conduction band could not be captured at
all because the deep level would already be occupied. Then recombination
5.3. Generation and annihilation of free charge carriers 519

Figure 5.3: Capture and emission of an electron and a hole at the same deep center.

Figure 5.4: Effects of deep levels for differing relations between the capture and
emission rates for electrons and holes.

would be impossible. Since the capture of an electron from the valence band
is equivalent to the emission of a hole to this band, the first condition to be
satisfied may be expressed as Cn >> Ep. On the other hand, the electron
captured from the conduction band must make the transition to the valence
band more quickly, or, in other words, the deep center must capture a hole
from the valence band more quickly, than the time it takes the electron cap-
tured by the deep center can be re-emitted back to the conduction band.
Thus C p >> En must hold. If different order of magnitude relations exist
between the four rates, the deep center can no longer function as a recom-
bination center, but has other effects (see Figure 5.4). We summarize these
in the following survey: The deep center works as
a) recombination center for Cn >> E p , C p >> En
b) electron capture center for C, >> Ep, C p << En
c) hole capture center for Cp>> En, C,<< Ep
d) generation center for Cp << En, Cn << Ep.
Verification of the three last cases, which were not explicitly discussed above,
520 Cbapter 5. Non-equilibrium processes in semiconductors

is left to the reader. With the background of these considerations, the unipo-
lar annihilation of charge carriers at deep centers appears in a new light. It
is a special case of the more general situation considered here, in which the
deep center can capture and emit both electrons and holes.
Thus far, recombination through deep centers has been considered only
qualitatively. To exhibit a quantitative treatment, we present equations
for the temporal rates of change of the electron concentration n, the hole
concentration p, and the occupation probability f of the deep center. These
equations are generalizations of the relat,ions which were already employed
above in describing the capture of carriers by deep centers. Here, they read

(5.50)

(g)Tmb(Z) - recomb +Nt (2) = 0. (5.52)

By means of relations (5.33) and (5.45) between, respectively, en, cn and ep,
cp, and using notation rno and rPo,introduced above, we have

(5.53)

1
= - [(l- f ) P t - fPl . (5.54)
(%)recomb TPo
Consider the case of stationary recombination. Then the occupation of the
deep level does not change in time, so that

(g) = 0, (5.55)

whence it follows from (5.52) that

(5.56)

i.e., the numbers of electrons and holes vanishing per volume and time unit,
are the same. Their common value is the number of electron-hole pairs
recombining per unit volume and time, so that the recombination rate R is

R=-($)
recomb
=-(a) recomb (5.57)

Equating (5.53) and (5.54), we have


5.3. Generation and annihilation of free charge carriers 52 1

from which R may be obtained as


n p - ntpt
R= (5.59)
+ + Tnoo) +P t )
~ p ~ ( nnt )

The relation
CI

ntpt = nopo = n:, (5.60)


permits R to be rewritten in the more compact form

with

as recombination coefficient. The expression (5.61) for the recombination


rate may be interpreted as follows: The driving force of recombination is the
deviation of the concentration product n p from its equilibrium value nope,
since the recombination rate is proportional to this deviation. The propor-
tionality factor T in general depends on the non-equilibrium concentrations
n and p . Only for small deviations from equilibrium is T approximately
constant. We will consider this case in more detail below.
Small deviations from equilibrium
Setting
n=no+An, p=po+Ap (5.63)
with [An1<< no and lApl p o , we first assume Ap = 0, supposing that the
non-equilibrium state has been created by unipolar electron excitation and
that it decays by recombination:

An#O, A p = O .
It follows that

An
R=- (5.64)
Tn
with
7, = no+ntTpO I Po+Pt,nO
~

(5.65)
PO PO
as the recombination lifetime of non-equilibrium electrons. In the case of
a n-type semiconductor, i.e. if p o << p t << no holds, then T n >> ~ p o , ~ n o ,
which means that the recombination lifetime is substantially longer than
the capture times for electrons and holes. This result is due to the fact
that the holes necessary for recombination of the non-equilibrium electrons
522 Chapter 5. Non-equilibrium processes in semiconductors

exist in the n-type semiconductor only in very low concentration, so that


the probability of recombination is relatively small and the recombination
lifetime is relatively large. In the case of a p-type semiconductor, i.e. for
no << nt << p o , then r, = rn0. This reflects the fact that the number of holes
is now large enough for every non-equilibrium electron to recombine. The
bottleneck for recombination is electron capture at the deep center.
For unipolar hole excitation, i.e. with

Ap#O, An=O,
one obtains analogous relations, and

(5.66)

with
rp =
no
~
+ nt rpo + PO-
+ pt TnO (5.67)
n0 n0
as the recombination lifetime of non-equilibrium holes. For an n-semiconduc-
tor one has rp = rPo,and for a p-type semiconductor rp >> rno, rpo.
The above results for the case of unipolar excitation may be summarized
by stating that the recombination lifetimes of minority carriers are substan-
tially smaller than those of majority carriers - provided the deviation from
equilibrium is weak.
Finally, we consider the case of bipolar excitation of non-equilibrium
carriers. In this case,

An = Ap.
The recombination rate is now given by

(5.68)

with
(5.69)

and
r = rpo(no nt) + + T,O(PO + P t ) (5.70)
no +PO
For an n-type semiconductor one has r = rpo, and for ap-type semiconductor
r = r,~. The recombination lifetime therefore equals the capture rate for
+
minority carriers. For an intrinsic semiconductor, we have no p o = Pni, so
that the recombination lifetime in this case is given by the expression

(5.71)
5.4. Drift current 523

Table 5.2: Recombination lifetimes q and intrinsic carrier concentrations ni for


several semiconductors. c ~ u i u=. T / V , denotes the capture cross section corre-
sponding t o ri and ni according to formula (5.71). T = 300 K . (Data compiled
from various sources.)

Material Gap r, (s) ni ozuiv.(cm2)


Si indirect 3.9 x lo4 1.5 x i o 1 O ioP2
Ge indirect 0.4 2.4 x 1013 5x
GaP indirect 3.4 x 10l2 2.7 5 x 10-21
GaAs direct 5.4 x lo2 1.8 x lo6 10-l~
InP direct 5.8 x 10 8.2 x lo6 10-15
InSb direct 7.3 x 10-7 1.6 x 1016 10-17

In Table 5.2 typical values of the lifetimes q for some semiconductors are
listed. It should be noted that there is a large variation of q , between
3.4 x 10l2 s for the indirect wide gap semiconductor GaP and 7.3 x s
for the direct small gap semiconductor InSb. For extrinsic semiconductors,
we have no + P O > 2 n i , so that T < q. This means that the recombination
lifetime through a particular deep center reaches its maximum value when
the semiconductor is in its intrinsic state. Doping reduces the recombination
lifetime of bipolarly excited non-equilibrium carriers.

5.4 Drift current


Consider a semiconductor in a spatially uniform electric field E. Apart from
its potential, which depends linearly on x,all other characteristic quantities
of the semiconductor, in particular, the concentrations n and p of free charge
carriers, are assumed to be spatially uniform. If the electric field vanishes,
the free charge carriers just perform thermal motions. The velocities of the
latter have the same probability in all directions in space, thus the average
momenta < m i x , > of electrons and < m;xp > of holes are zero ( m i and
m; denote the effective masses of, respectively, electrons and holes). In the
following treatment of charge carrier drift, we will confine our considerations
to electrons as free carriers. The results can be transferred to holes without
any complications, involving only notational changes.
In the presence of the electric field, the electrons are accelerated and
the average value of electron momentum < mixn > has a non-vanishing
524 Chapter 5. Non-equilibrium processes in semiconductors

component in the direction of the field. Its rate of change [$ < m i i n > field
1
per second, according to equation (3.110), equals the force impressed by the
field on a Bloch electron. whence

[& < mkkn >


1 field
= -eE. (5.72)

In addition to the electric field there are frictional forces acting on the Bloch
electron which arise from the semiconductor itself. The interactions of an
electron with lattice vibrations and impurity atoms or defects result in scat-
terings which act to brake its acceleration along E, producing a frictional
effect similar to that in a viscous medium. Often, this frictional resistive
force grows stronger in proportion to the average electron speed, so that one
may put the corresponding rate [$ < mkxn >Iez of momentum change per
second due to collisions with phonons and crystal imperfections, proportional
to the average momentum, whence

(5.73)

with l / r i as proportionality factor. If there is a momentum excess at initial


time t = 0, < m;xn(0) >, the solution of this equation reads

< mkxn(t) > = < rn:jcn(o) > e-t/T:. (5.74)


Thus, the momentum excess decays exponentially, relaxing to zero, as one
may say. In this context, one terms r;E as the momentum relaxation time.
This time should not be confused with the previously introduced electron
lifetime 7, for capture or recombination. The two characteristic times de-
scribe completely different physical phenomena. In general the momentum
relaxation time is shorter than the capture and recombination lifetimes.
The total-momentum balance is governed by Newtons equation of mo-
tion,

whence

d 1
- < mkxn > = -eE- - < mn
* kn >. (5.76)
dt r
;
For steady state dc current flow, the left hand side of this equation vanishes
and the concomitant constant value of the average velocity is
5.4. Drift current 525

(5.77)

which is referred to as the drifi velocity.


For a field strength of 1 Vcm- the drift velocity has absolute value

(5.78)

This quantity is a measure of the mobility of an electron per unit applied


force, therefore pn is simply referred to as electron mobility. Using equation
(5.78), the drift velocity of (5.77) may be written in the form

< X, > = pnE. (5.79)


The average electron current density j,, defined as

j, = - e n < kn >, (5.80)


may be re-expressed in terms of p, as

j , = enp,E. (5.81)
For the electron conductivity D,, as it occurs in Ohms law, one t,hus obtains

on = enp,. (5.82)
This conductivity expression will be used frequently below. It can be trans-
ferred to holes without difficulty, reading

Tp = eppp, (5.83)
where the hole mobility p p is given by

er;
Pp= - - y I (5.84)
mP

with r; signifying the momentum relaxation time of holes.


Since electrons and holes in a semiconductor are always present simultane-
ously, the total current density is the sum of the electron and hole current
densities,

j = .in +j p , (5.85)
and the total conductivity D is the sum of the two partial conductivities,

T = un
C + up. (5.86)
526 Chapter 5. Non-equilibrium processes in semiconductors

Table 5.3: Electron and hole mobilities for pure semiconductor materials at T =
300 K . (After Landoldt-Bornstein, 1982.)

1000
Si

200
40
N=10"
100
20
Temperature ("C) I O U
-50 0 50 100 150 200
Temperature ("C 1

Figure 5.5: Temperature dependence of the electron and hole mobilities of Si for
various doping concentrations.

For extrinsic semiconductors of n- or p-type, one of the two charge carrier


concentrations exceeds the other by many orders of magnitude. To a good
approximation, the total current density for an n-type semiconductor can
be identified with j,, and for a p-type semiconductor with j,.
We will tabulate some values of the transport parameters above, and
discuss aspects of their characteristic dependencies. Since the charge carrier
concentration of a given semiconductor can vary over a wide range, it is more
meaningful to discuss p n and p p rather than of u, and up. The mobility unit
one commonly uses is cm2V-ls-'. If the mobility is unity in this unit, then
a field of 1 Vcm-' causes a drift velocity of

1 cm2v-1s-l x 1 v c m - l = 1 cms-'. (5.87)

In Table 5.3 the mobilities of electrons and holes are given for some important
semiconductors at room temperature. That the mobilities of electrons exceed
those of holes, is to be attributed to the smaller effective masses of the
electrons in these materials.
The mobility depends on the interaction of the carriers with phonons through
5.5. Diffusion and annihilation of free carriers 52 7

Figure 5.6: Dependence of the


electron and hole mobilities on 5 loT- 3004
K 7
doping concentration for Ge, Si
N=- 103
and GaAs at T = 300 K . (After
Sze, 1981.)
-5
,102

Doping concentration ( ~ r n - ~ )

the momentum relaxation time, and the same may be said for temperature
dependence. In Figure 5.5 this dependence is shown for Si. The decrease
of mobility with rising temperature is to be attributed to the growth of
the amplitudes of the lattice oscillations or, in other terms, of the num-
ber of phonons. In extrinsic semiconductors the mobility also depends on
the doping concentration, because doping atoms constitute collision centers.
Generally, the momentum relaxation times, and through them also the mo-
bilities pn and p p , decrease with rising doping concentrations. Examples of
such dependencies are shown in Figure 5.6.

5.5 Diffusion and annihilation of free carriers


In this section, we consider the concentrations n and p of electrons and holes
to be spatially inhomogeneous. A direct consequence of this assumption is
that the total charge density, which consists of the charge density of the
free carriers and that of the ionized doping atoms, is no longer zero locally,
although the total charge vanishes upon integration over the entire sample.
The non-vanishing local charge distribution is accompanied by spatially in-
homogeneous electric fields. Later, we will take account of these fields, but
for the moment we omit them. To be consistent, we correspondingly ignore
the charges carried by electrons and holes.
Spatial inhomogeneities of the electron and hole concentrations n ( x ) and
528 Chapter 5. Non-equilibrium processes in semiconductors

p(x) imply the existence of non-vanishing gradients O n ( x ) and Vp(x). The


latter cause spatially inhomogeneous diffusion currents in and ip of, respec-
tively, electrons and holes. Considering first the case of electrons, the diffu-
sion current is given by

in = -DnVn. (5.88)
Because of its inhomogeneity, this current leads to temporal changes of the
local electron concentration, whereby the latter may take values which de-
viate from local equilibrium. As a result, generation and annihilation pro-
cesses will take place which tend to restore local equilibrium values. The
overall balance of the electron concentration change is described by the con-
tinuity equation (5.20). The right-hand side of this equation is specified
such that generation processes are omitted while capture and emission pro-
cesses at deep centers are taken into account. Then the annihilation term
- ( d ~ ~ / d t ) ~ ~ equals
~ i h i l (dn/dt),,pture. For small deviations from equilib-
rium, (an/dt)capture may be identified with expression (5.39), whence

dn n - no
-+V.i n- (5.89)
at 7,

For the sake of simplicity, we suppose that the electron concentration n ( x )


changes only in x-direction. Then only the x-component of in is non-zero and
we denote it by in. In the continuity equation, we eliminate V.in = (&,/ax)
by means of the diffusion equation (5.88), with the result

-d -nD d2n n-no


(5.90)
at nax2 - rn .
For steady state this yields

(5.91)

The determination of a unique solution of this equation requires specification


of boundary conditions, which we choose as follows: at x = 0 a particular
non-equilibrium value

n(x = 0) = no + An0 (5.92)


of the electron concentration is maintained, and n ( x ) will decay to the equi-
librium value no at x = 00. In the interval 0 < x < 00, the solution is
written in the form

(5.93)
5.5. Diffusion and annihilation of free carriers 529

Figure 5.7: Profile of the non-


equilibrium electron concentra-
tion as a function of position
x in the presence of diffusion
t
and charge carrier annihilation
(schematically).

0
X-

Then it is readily seen that

n(2)= no + Ange-"ILn, (5.94)

where
L,= &. (5.95)

The solution (5.94) admits the following simple physical interpretation (see
also Figure 5.7). Diffusion tends to disassemble existing concentration gra-
dients, transferring part of the distribution near z = 0 further to the right.
However, since the lifetime of non-equilibrium charge carriers is limited, this
process cannot be fully accomplished, as electrons are captured during the
transfer process, leading to the exponentially decaying profile of the electron
concentration given in equation (5.94). The characteristic decay length L ,
can be interpreted as the distance to which an electron diffuses, on the aver-
age, before it is captured. One refers to L , as the dzfluszon length. L , is the
larger for larger lifetime 7,. One can illustrate the interpretation above by
means of a fluid current moving over porous ground. At a particular point
x = 0 the fluid current level is maintained at a fked height. In the direction
of the current flow, the fluid level drops because fluid is absorbed into the
porous ground. Sufficiently far downstream, practically no fluid will arrive
because most of the h i d was absorbed into the ground further upstream.
Nevertheless, one has a continuous fluid current at x = 0, which provides
exactly as much fluid as is absorbed along the entire pat,h between x = 0
and z = 00. The analogy to the diffusion of electrons with finite lifetime is
obvious. Furthermore, the current flow at z = 0, in this case i,(z = 0), may
be determined from (5.88) and (5.94) as

(5.96)

These considerations for electrons can be immediately transferred to holes.


In analogy to equation (5.91) one obtains
530 Chapter 5. Non-equilibrium processes in semiconductors

Table 5.4: Diffusion coefficients and diffusion lengths for several pure semiconduc-
tors at T = 300 K . Diffusion coefficients are calculated by means of mobility data
of Table 5.3 using Einstein's relation.

Si
Ge 100.8
GaAs 227.5 10.3 10-8 3.2

(5.97)

For boundary conditions analogous to those in equation (5.92), we have

with
(5.99)

as the diffusion length of holes.


In Table 5.4, diffusion lengths of electrons and holes are listed for several
semiconductor materials. The lifetimes r, and rp depend strongly on the
degrees of purity and crystallographic perfection of the semiconductors. In
Table 5.4 a value of s was assumed for T,,~. However, in very pure
and perfect materials, considerably larger values are possible. The diffusion
lengths are then much larger than those given in Table 5.4, and they can
reach hundreds of microns.

5.6 Equilibrium of free carriers in inhomogeneous-


ly doped semiconductors
In this section we consider semiconductors with spatially inhomogeneous
doping, and explore the thermodynamic equilibrium state of such a spa-
tially inhomogeneous system. The more general case, which includes semi-
conductor materials having an inhomogeneous chemical composition, will
he discussed in Chapter 6. As before, we can consider the semiconductor to
be composed of a subsystem of atomic cores and another of the free charge
carriers. Both subsystems are taken to he spatially inhomogeneous here.
5.6. Equilibrium of free carriers in inhomogeneously doped semiconductors 531

Complete equilibrium of the total system can only exist then if the inho-
mogeneities have equalized out in both partial systems. This means also
that the doping of the semiconductor material must have become spatially
homogeneous. The process which brings this about, the diffusion of dopant
atoms, is so very slow, however, as compared to the carrier processes, that
one can completely neglect it. This means that, the subsystem of atomic
cores effectively remains in the non-equilibrium state in which it started;
this state is frozen, so to say. The subsystem of the free charge carriers,
however, will relatively quickly pass into a new stationary state. On the one
hand, this state will differ from the equilibrium state of free charge carriers
in a homogeneous semiconductor - the concentrations n and p of the elec-
trons and holes will depend on position coordinate x,as a consequence of the
inhomogeneity of the dopant concentrations. On the other hand, all current
densities of the free charge carriers will also vanish in the inhomogeneous
case, just as in the homogeneous case. A stationary final state conforming
to this characterization may be understood as a thermodynamic equilibrium
state of the free charge carriers of the spatially inhomogeneous semiconduc-
tor. The nature of thermodynamic equilibrium in this sense will be discussed
below.
We consider only spatial inhomogeneities in the x-direction. The com-
ponents j , and j , of the electric current densities of electrons and holes in
this direction are given by the phenomenological equations

(5.100)

dP
, = a,E(x)
j - eDp-. (5.101)
dx
Here E(x) represents the electric field that occurs in consequence of the
spatially inhomogeneous charge carrier distributions, initially omitted in the
previous section. In equilibrium the current densities of electrons and holes
must vanish. That this must hold for the electron and hole currents sepa-
rately, and not just for the sum, follows from Boltzmanns equation which
states that in equilibrium the partial currents of the subsystems of electrons
having a fixed energy must vanish separately (not just after summing over
all energy values). The holes are, of course, also a subsystem of the total
electron system with particular energy values, namely the energies of the
unoccupied states of the valence band. Hence, in equilibrium we have

+ dn
0 = ~ , E ( x ) eDn--,
dx
(5.102)

0 = apE(x) - eDp-.dP (5.103)


dx
532 Chapter 5. Non-equilibrium processes in semiconductors

The two electron current contributions in (5.102), the drift current v n E and
the diffusion current eD,(dn/dx), need not vanish separately. The same
holds for the hole current contributions a p E and - e D p ( d p / d x ) in equation
(5.103). Only the sum of these contributions must be zero. Thus we arrive
at the conclusion that in thermodynamic equilibrium the drift and diffusion
currents of each type of carrier must exactly compensate each other.
Our object now is to transform this observation into a general condition
for thermodynamic equilibrium in systems with spatially inhomogeneous
doping. To this end we express the field strength E of equations (5.102)
and (5.103) in terms of the electrostatic potential cp, and replace c n and up,
respectively, by the right hand sides of equations (5.78) and (5.83). Thus,
we obtain

dYJ
o=- ~ n -
dx
+ Dn-ddx [Inn ( x ) ], (5.104)

dcp d
O = -pp- - Dp- [ h p ( ~ ) ] (5.105)
dx dx
In the case of local equilibrium, the charge carrier concentrations n(x) and
p(x) depend on the chemical potential p(x) in just the same way that n and
p in expression (4.59) depend on E F ,

(5.106)
Substituting these expressions in (5.104) and (5.105) we have

(5.107)

(5.108)

In order to satisfy these two relations simultaneously, it is necessary that

(5.109)

Equations (5.107) and (5.108) will now be applied to a particular spatially


inhomogeneous system for which the space variation of the chemical potential
is known. This consideration will result in the Einstein relation.

Einstein relation
Because the chemical potential for holes is the same as that for electrons,
we can restrict ourselves to the latter. We assume that at a given position
z = xo there is an infinitely high potential wall for these particles (see Figure
5.8). Electrons can reside on the right side of the wall, but not on the left, as
5.6. Equilibrium of free carriers in inhomogeneously doped semiconductors 533

Figure 5.8: Illustration of the


derivation of the Einstein rela-
. .
tion.
-
-
X
4
aJ
I
01
w

I
I

xo
X

they are reflected by the wall if they move toward the left. The situation is
comparable with that of molecules of air above the surface of the earth. As
in the latter case, in which the equilibrium vertical distribution of particles
is given by the law of atmospheres, we have here

Comparing equations (5.110) and (5.106) it follows p(z) = ecp(z). Since


equation (5.107) applies also in this special case, the factor multiplying
( d p l d z ) must necessarily be ( l / e ) , whence we conclude that

(5.111)

This equation is the Einstein relation It holds for holes as well as electrons,
kT
D, = - p P . (5.112)

A generalization of Einsteins relation for degenerate charge carrier gases is


possible, but we will not discuss this here.

Condition for thermodynamic equilibrium of free carriers

Combining equations (5.111) and (5.107), we obtain the desired general con-
dition for thermodynamic equilibrium of free charge carriers,

9-e d- v = 0 (5.113)
dx dx
or
p ( z ) - ecp(z) = const (5.114)
The quantity on the left hand side of this equation is referred to as the
electrochemical potential. In the case of vanishing electric potential, the
electrochemical potential is just the same as the simple chemical potential.
This also remains true if the electric potential has a non-zero but spatially
534 Chapter 5. Non-equilibrium processes in semiconductors

constant value which, as we know, is irrelevant. The terms Fermi energy


or Fermi level which previously, in the absence of spatially varying electric
potential, were used for the simple chemical potential, will now be used
for the electrochemical potential. With this terminology, relation (5.114)
becomes

(5.115)
In thermodynamic equilibrium, the electrochemical potential of the free
charge carriers is spatially constant, and the constant value is equal to the
Fermi energy. The chemical potential and the electric potential themselves
can change, however. It is appropriate to bear in mind here that the ini-
tially assumed spatial inhomogeneity of the semiconductor is based on the
inhomogeneity of its doping. The formal expression of this inhomogeneity is
the spatial variation of the chemical potential. To properly adjust the ther-
modynamic equilibrium state, the spatial variation of the electric potential
must just compensate the spatial variation of the chemical potential.
The equilibrium state of equation (5.114) represents an equilibrium in
both the local and in the global senses. That in the local sense was pre-
supposed (approximately) at the outset with the use of phenomenological
equations for the current densities of the charge carriers. Moreover, the equi-
librium state described by (5.114) represents a relative equilibrium, because
it was derived under the assumption that the initial spatial inhomogeneity
of doping, and thus the inhomogeneity of the chemical potential, remains
unchanged. Over very long periods of time, this is assumption ceases to be
valid. However, this does not affect the equilibrium condition because it is
only essential that the inhomogeneity of doping does not change during the
relaxation of the charge carriers towards equilibrium, a condition which is
always fulfilled.
As one may expect, the equilibrium condition (5.114) or (5.115) will
play a central role in the treatment of semiconductor junctions in the next
chapter.
535

Chapter 6

Semiconductor junctions in
thermodynamic equilibrium

The distinguished role of semiconductors in the modern world stems from


devices which can be fabricated from these materials and, like transistors
and semiconductor lasers, have opened the door to the present computer
and communication technologies. The operation of semiconductor devices
depends largely on effects which occur at semiconductor junctions, in ad-
dition to the properties of homogeneous semiconductors which have been
treated almost exclusively in this book thus far. A semiconductor junction
is understood as a structure composed of two interconnected layers, of which
at least one is a semiconductor material. Various semiconductor junctions
are shown in Figure 6.1 schematically. In case (a) the two layers are made
of the same semiconductor material, e.g. silicon, but one of the layers is
p-doped and the other n-doped. This is called a pn-junction In case (b)
two layers of different semiconductor materials are connected. The doping
type may, but need not be different. This structure is called a semiconductor
heterojunction. We encountered such semiconductor heterostructures above
in section 3.7. If a semiconductor layer is combined with a metal layer,
then one has a metal-semiconductor junction, and if it is connected with
an insulator layer it is an insulator-semiconductor junction Formally, one
can also include semiconductor surfaces here, which correspond to vacuum-
semiconductor junctions.
Each of the semiconductor junctions identified above is associated with
an important physical discovery. Several of these discoveries were even hon-
ored with Nobel prizes. Moreover, each of these junctions plays a role in
electronic devices, and in most cases a semiconductor junction is crucial
for operation of the device. Electric rectification at a metal-semiconductor
junction, the so-called Schottky contact, was historically the first semicon-
ductor junction to be used technologically. In this particular case, funda-
536 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

a)

n -type

b)

cnSemic. Metal

Figure 6.1: The various semiconductorjunctions.

mental general laws of semiconductor junctions were first recognized, par-


ticularly by Schottky and Mott. It is thus astonishing that, in this case of
the metal-semiconductor junction, a complete microscopic understanding is
elusive even now. The pn-junction, more precisely the combination of a pn-
and a np-junction, underlies the bipolar transistor, for which the inventors
Shockley, Bardeen and Brattain received the Nobel prize for physics in 1956.
Also, the tunnel diode, for which the Japanese physicist Esaki earned the
Nobel prize in 1973, rests on the pn-junction. This device employs electron
tunneling between the valence and conduction bands of such a junction. The
unipolar field effect transistor (MOSFET) contains, beside a pn-junction,
also an insulator-semiconductor junction as an active structural element. In
experimental studies of the 2-dimensional electron gas in specially tailored
silicon MOSFET's, the quantized Hall effect was discovered by von Klitz-
ing. This accomplishment was honored with the Nobel prize for physics in
1985. The 2-dimensional electron gas and the quantized Hall effect may also
be observed in semiconductor heterojunctions. Combining heterojunctions
with pn-junctions, efficient light emitting diodes (LEDs) and semiconductor
injection lasers may be fabricated. Heterojunctions also form the key struc-
tural elements of which the superlattices and quantum wells of section 3.7 are
composed. The advent of these systems has opened a completely new and
fruitful area in semiconductor physics and electronics, that of semiconductor
microstructures.
In general terms, the operation of electronic devices based on semicon-
ductors may be described as follows: The thermodynamic equilibrium of a
semiconductor junction is disturbed by an external perturbation, for exam-
6.1. pn-junction 537

ple, by applying a voltage or by exposing the sample to light. The result


of the perturbation are large changes of some properties of the junction,
for example, its electric resistivity. These changes are the effects employed
to advantage in semiconductor devices. To understand the perturbation of
an equilibrium state of a semiconductor junction, one must first deal with
the equilibrium state itself. This will be done in the present chapter. Non-
equilibrium processes will be treated in Chapter 7. To initiate discussion of
the equilibrium state, we start with the pn-junction.

6.1 pn-junction

pn-junctions are fabricated such that either the surface region of a homoge-
neously doped n (or p)-type sample, is p (or n)-doped by means of diffusion
or ion implantation, or that a p (or n)-type layer is deposited on a n (or
p)-type substrate by means of epitaxy. In the particular case of Si, neutron
bombardment may also be used. Due to the nuclear reaction ;gSi n + ---f

+
:aSi y -+ + p, a previously p-doped sample becomes n-doped in the
surface region exposed to the neutrons.
In the following theoretical description of the pn-junction we assume
homogeneously doped p - and n-regions left and right of the plane normal
to the z-axis of a Cartesian coordinate system (see Figure 6.2a). The p -
region extends to -00 and the n-region to +co. The transition from the p -
to the n-region takes place abruptly at z = 0. We refer to this position as
the nominal transition. In the transverse plane normal to the z-axis, the
two semi-infinite semiconductor samples are taken to extend to infinity in
all directions. Thus, there are no dependencies on y and z , provided the
semiconductor materials are homogeneous in the transverse plane, which we
take to be the case. With this, the problem is effectively one dimensional.
The carrier concentration distribution along the x-axis before equilibrium is
established is shown schematically in Figure 6.2a. Conceptually, this corre-
sponds to the instant at which the junction is formed, so that the carrier
distribution has not yet had enough time to adjust to form a new equilibrium
state. Considering that the p-region also has electrons which are minority
carriers in it, and that the n-region also has holes as minority carriers there,
one must distinguish the region in which these concentrations are meant.
We indicate this by affixing a subscript to n and p . The concentrations in
the n-region will be denoted by n, and p,, and those in the p-region by np
and p p . Thus, n,, and p p are majority carrier concentrations, while np and
p, are minority carrier concentrations. The different values of the charge
carrier concentrations in the n- and p-regions are related to different values
E F , and E F of ~ the bulk Fermi energies in these regions, such that
538 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

1---
I

D i f f h o n tc Drift

-ecp (x)

I I

C)

I
I
Recombi-
nation I
I
I
I
I
1 I
I

I
X

I I

Figure 6.2: Non-equilibrium state of a pn-junction (a), from which the establish-
ment of thermodynamic equilibrium develops (b,c).
6.1. pn-junction 539

6.1.1 Establishment of thermodynamic equilibrium


The situation shown in Figure 6.2a certainly does not represent an equilib-
rium state. The charge carrier concentrations are spatially varying functions
n(x) and p ( z ) , and because np << n, and p, << pp, strong gradients of n(x)
and p(x) exist in the vicinity of the nominal transition x = 0. They cause
electrons to diffuse to the left into the p-region, and holes to the right into
the n-region (Figure 6.2b). Consequently, the n-region becomes positively
charged, and the p-region negatively. In the terminology of Chapter 5 , one
can say that a space charge region is formed. It has an associated electric
field E ( z ) and an electric potential cp(z). The dependence of these quantities
on x is shown in Figure 6.2b (right column), illustrating results which will
actually be obtained later by detailed analysis. For our immediate consid-
erations, only their qualitative behavior is important. In Figure 6.2b the
band edges are shown as position-dependent quantities. This representa-
tion is surprising initially, since the band edges are energy eigenvalues of
the Schrodinger equation, and as such they do not depend on space coor-
dinates. Such a representation in terms of position-dependent band edges
is nevertheless meaningful in the present case, based on the fact that cp(z)
is a smoothly varying potential in the sense of section 3.3. Therefore, the
effective mass equation can be applied. For the envelope function Xc(x) of
the conduction band electrons this equation reads

and for the envelope function Xv(x) of the valence band holes it has the
form

According to equation (6.3), the electrons at the conduction band edge ex-
hibit a potential energy E, - ep(z), and the holes at the valence band edge
exhibit a potential energy ecp(z). These energy functions are depicted in
Figure 6.2b. In this, one has to recognize that the energy e p ( z ) of the holes
counts negatively, so that on the positive electron energy scale one must
therefore plot -ep (z).
The electric field in the space charge region drives a current in the direction
opposite to that of the diffusion current (Figure 6.2b). The concentrations
540 Chapter 6 . Semiconductor junctions in thermodynamic equilibrium

of carriers which diffuse from the region where they are majority carriers
into the regions where they are minority carriers, are much larger than the
equilibrium minority concentrations there. Thus they will recombine with
the abundantly available majority carriers in these regions (Figure 6 . 2 ~ ) In .
this process, the concentrations of both types of free charge carriers decrease
in the space charge region, as shown in Figure 5.2. The electric charge bal-
ance as well as the electric field and potential distributions are, however,
not influenced by such recombination since the annihilation of electrons and
holes takes place simultaneously, and in the same measure. Overall, a p n -
junction, in the time interval immediately following its formation, undergoes
a relatively complicated interplay between various non-equilibrium processes
- the diffusion and drift of majority carriers, the formation of a space charge

region and of related electric field and potential distributions, as well as re-
combination of non-equilibrium minority carriers with equilibrium majority
carriers.
The processes under discussion act in such a way that thepn-junction ap-
proaches thermodynamic equilibrium. The state to which the system relaxes
is, however, not a static but a dynamic equilibrium state. Of course, it has no
net current, but only because the drift and diffusion currents fully compen-
sate each other. In regard to the equilibrium state which finally emerges, the
details of the processes which establish equilibrium do not matter, what is
important is only that there are such processes. Independently of the nature
of these processes, specific local values of the chemical potential p ( z ) , electric
potential p(z), and charge carrier concentrations n ( z ) ,p(z) are established
in equilibrium. The latter are given by the relations (5.106) discussed in
Chapter 5, and we display them here once again:

Far from the nominal pn-transition at z = 0, recombination processes cause


the concentrations n ( z ) and p ( z ) to take their equilibrium values, whence

n(-m) = np, n(+m)= nn, (6.6)


~ ( - 0 0 )= ~ p , P(+w)= Pn. (6.7)
Considering (6.2), we have

P ( - m ) = EFp, P ( + m ) = EFn. (6.8)


As discussed in Chapter 5, thermodynamic equilibrium is characterized by a
spatially constant electrochemical potential ~ ( z-) ep(z). The calculation of
the electric potential p(z), is carried out using the Poisson equation (5.19).
In this, one only needs to specify the charge density p ( z ) on the right-hand
side in accordance with the relationships which hold for the pn-junction.
6.1. pn-junction 541

Before doing this and solving the Poisson equation, we will first determine a
relatively easily accessible integral quantity, namely the potential difference
U D , which exists between the left and the right end of the pn-junction in
consequence of the charging of the two regions. This potential difference is
called &;fusion voltage, because diffusion is the agent actually responsible for
its existence. The diffusion voltage provides one of the boundary conditions
needed for the solution of the Poisson equation.

6.1.2 Diffusion voltage


The diffusion voltage
UD cP(m) - cP(-m)
can be determined by writing down the electrochemical potential p ( z ) -
ecp(z) once at z = -m and again at z = +m, and requiring the two values
to be equal, whence

p(-m) - ecp(-oo) = p ( + w ) - ecp(+m). (6.10)


Using (6.8) and (6.9), we have
1
UD - [ E F ~- E ~ p l . (6.11)

The Fermi levels may be expressed in terms of the majority carrier con-
centrations nn and p, employing equations (6.5) to (6.8). First of all, we
have

nnPp - N ~ N ,-Eg/kTe ( E F-~E ~ p l / k T . (6.12)


Introducing ni from formula (4.86) and taking the logarithm, it follows that

(6.13)

To provide an estimate of the order of magnitude of the diffusion voltage,


we proceed as follows: Let be T = 300 K , i.e. k T M 25 m e V . With
n, L- p, = 10I6 cm-3 and ni = 10" ~ r n -(roughly
~ corresponding to the
intrinsic concentration of Si), we obtain U D = 0.7 e V . A comparable value
follows if one assumes that the Fermi level in the n-type semiconductor
lies just below the conduction band edge, and in the p-semiconductor just
above the valence band edge. Correspondingly, the difference of the two
Fermi energies, which equals the digusion voltage, takes a value which is
somewhat smaller than the energy gap. Thus, under realistic assumptions,
we find that diffusion voltages have substantial magnitude.
Diffusion voltages cannot be measured or used, however, in any way
involving interconnection of the two ends of the p- and n-regions by wires
542 Chapter 6. Semiconductorjunctions in thermodynamic equilibrium

+- - + bl
+- P I:
-+
n ;:

I Metal I -- ++

I 1
+
+
+

Figure 6.3: Measuring the diffusion voltage of a pn-junction.

(Figure 6.3a), or by bringing the two ends directly in contact with each other,
bending the p - and n-regions (Figure 6.3b). No current will flow through
closed circuits obtained in this way. The total voltage drop in such circuits
is zero, since, beside the original pn-junction already present a second one is
created by closing the circuit, acting in the opposite direction. The metallic
connections in Figure 6.3a only play a passive role and do not affect the
voltage balance of the circuit - the sum of the diffusion voltages of the two
metal-semiconductor junctions exactly equals the negative of the diffusion
voltage of the pn-junction. Nevertheless, the diffusion voltage can actually
be measured, namely, in an electrostatic experiment. This can be done, for
example, when the two ends of the p- and n-regions are bent to be close
to each other, leaving only a small gap between them (Figure 6 . 3 ~ ) .In the
gap there is an electric field associated with the diffusion voltage, whose
existence may be demonstrated by means of a small test charge.

6.1.3 Spatial variation of the electric and chemical potentials:


Schottky approximation
The electric potential p(z) obeys the Poisson equation (5.19) with p ( z ) taken
from (5.16). A particularly important feature of this equation is that its
charge sources on the right-hand side themselves depend on the potential
p(z). We consider this dependence initially for the free charge carrier part
of the charge sources. One obtains this from equations (6.5) for n ( z ) and
p ( z ) , expressing the chemical potential p ( z ) in terms of the electric potential
p(z) using the relation
6.1. pn-junction 543

whence

(6.16)
and with 6.13) we find
2
n ( x ) p ( z ) = ni (6.17)
The mass action law, which was written down for spatially homogeneous
semiconductors in equation (4.85), therefore also holds locally for inhomo-
geneous semiconductors. Such a result is not surprising since this law is
predicated on the existence of thermodynamic equilibrium, and local equi-
librium was assumed in the inhomogeneous case considered here.
The concentrations of ionized donors NDf and acceptors N Z depend on
the potential (p(z),at least in the general case. In a frequently occurring
special case, namely that of complete ionization, this dependence no longer
exists. We will assume that this case applies, so that
) N&),
~ o f ( x= N A ( X )= N A ( Z ) . (6.18)
Since donor atoms are present only in the n-region, and acceptor atoms
only in the p-region, and since they are homogeneously distributed there in
accordance with our initial assumption, it follows that

(6.19)

0, x > 0,
N ~ ( X=) (6.20)
N A , x < 0.
The Poisson equation for cp(x) thus reads

(6.21)

We seek a solution of this equation which accounts for the existence of the
diffusion voltage (6.9). In order to get a unique solution one more condition
is needed. This may be obtained from the observation that far from the
nominal transition at x = 0, everything should behave as it would in an in-
finite bulk sample. This means, in particular, that the electric field strength
has to vanish at x = f c o , whence

(6.22)
544 Chapter 6. Semiconductorjunctions in thermodynamic equilibrium

Figure 6.4: Illustration of the Schottky approximation.

The first part of this relation holds because of total charge neutrality of
the pn-junction. It is not an independent condition, and does not provide
additional information.
The scalar potential p(z) is determined only up to an additive constant.
Without any loss of generality we may chose this constant such that

(6.23)

The explicit determination of the solution of the Poisson equation (6.21) is


a difficult task, mainly because the function p(z) to be calculated enters the
right-hand side of this equation via the charge carrier concentrations n ( z )
and p ( z ) as given by (6.15) in an extremely non-linear way, namely expo-
nentially. It is therefore necessary to resort to approximations or numerical
calculations.
Here, we will discuss the former. First of all, one must consider which
approximations are suitable under existing conditions. The answer may be
obtained from a qualitative estimate of the potential profile p(z) illustrated
in Figure 6.4. The result of the estimate has already been used in Figure
6.2 to construct -ep(z). Far from the nominal pn-transition at z = 0 there
is local charge neutrality, as for a homogeneous semiconductor sample. In
consequence of the Poisson equation and the boundary condition (6.22), the
potential will approximately retain the values p(-m) and p(+oo) which it
has, respectively, at z = --oo and z = +m, in approaching the position
x = 0 from --oo and +m. In the vicinity of the nominal transition at z = 0
a space charge region is formed, as outlined above, and it is here that the
6.1. pn-junction 545

n= N,
I Depletion region
---- I+++,+
Bulk region Ispace charge region1
I
Bulk region

Figure 6.5: Space charge region, depletion region and bulk region of a pn-junction.

potential will undergo significant variation. We will denote by x , the first


position in the n-region where, in the approach from +CQ, space charge starts
to develop substantially. Correspondingly, the potential starts to undergo
major change at x,, where it is approximately given by cp(x,) = cp(00).
Since the total change of cp(x) over the space charge region amounts to U D ,
and since U D is, in general, substantially larger than k T , the point is quickly
reached where cp(x) has fallen off to the value cp(x,) - k T / e . We denote this
point by x n - A x , , so that

V ( x n - A X , ) = cp(xn) - k T / e . (6.24)
The electron concentration n ( x ) of (6.15) at 2 , - A x , has already decreased
down to the e-th part of its equilibrium value nn (see Figure 6.4). Since it
decreases further as x decreases, one can approximately set it to zero for
x , - A x , , whence

0, x <~ n A-x , ,
n(x)= (6.25)
n,, x > x,.
An analogous conclusion can be drawn for the hole concentration p(x). Us-
ing similar notations x p and A x p for holes as were used for electrons, we
approximately have

(6.26)

If it turns out, as is actually the case, that A x , << x , and A x p <<I x p I hold,
then we can make the further approximation of taking the two intervals A x ,
and Axp to be vanishingly small. We will do this, understanding this change
to be made in equations (6.25), (6.26) without rewriting these formulas.
The simplification described in the resulting equations (6.25) and (6.26) is
called the Schottky a p p r o x i m a t i o n In the pn-junction, we can now identify
four regions in which the total charge density has constant values, with
jumps from region to region (see Figure 6.5). For x < x p and x > x , the
546 Chapter 6 . Semiconductor junctions in thermodynamic equilibrium

charge density is zero, because the charge of the ionized impurity atoms is
compensated by the charge of the free electrons and holes. Between x p and
I , , free charge carriers are absent, so for x p < x < 0 we have the charge
density - e N A of the ionized acceptors, and for 0 < x < x , the charge density
e N D of the ionized donors. Because of the absence of free charge carriers
between x p and 0 and between 0 and x,, these two regions are called depletion
regions. Jointly, the two regions together form the space charge region, The
adjacent regions on the left and right hand sides are called bulk regions. The
Poisson equation associated with the charge distribution described above
reads

d2P
--
dx2 - {I -No,
NA,
0,
O < X <xn,
xp
x
< x < 0,
< xp.
(6.27)

The boundaries x , and x p of the space charge regions are undetermined


so far. One relation between them follows from the total charge neutrality
condition,
X,ND X ~ N=+ A 0. (6.28)
It states that the total positive charge of the n-side of the space charge region
equals the total negative charge of the pside of this region. A second relation
for the determination of x , and x p follows from the boundary condition
(6.22), which will be used later.
In solving the Poisson equation (6.27), attention must be given to the
fact that, notwithstanding discontinuity of the right-hand side at x = 0, the
only admissible potentials p(z) are those which are continuous everywhere
together with their first derivatives. The solution of the boundary value
problem (6.27), (6.9) and (6.22) which meets this requirement is given by

(6.29)

x < xp.
At the position x = 0, cp(x) is automatically continuous and vanishes, as it
should according to equation (6.23). The continuity of the h t derivative
at x = 0 follows from the charge neutrality condition (6.28). The diffusion
voltage condition (6.9) yields

(6.30)
6.1. pn-junction 547

Introducing the abbreviation w = I, - x p for the width of the space charge


region, it follows from (6.28) that

NA ND
(6.31)
2, =
ND + N Aw , xp=-
+N A ~ .
Applying equation (6.30) we get the result for w as

(6.32)

The potential values at x, and x p are given by

The calculated potential profile cp(x) corresponds in every detail to that


which we assumed qualitatively in motivating the Schottky approximation.
We also can now specify the validity limits of this approximation quanti-
tatively. In fact, substituting (6.29) for cp(z) into the condition (6.24), we
obtain

(6.34)

This length, which (apart from a factor a)


is the Debye screening length
for the potential U D in the n-region, must be small in comparison with the
width x, of the n-part of the space charge region, in order for the Schottky
approximation to be valid. A corresponding relation must be satisfied for
the p-region. With the help of equations (6.31) and (6.32) these conditions
may be written in the form

fi< d-m, n - region,

f i <J-J~TT;;,
< p-region. (6.35)
In order for these relations to be fulfilled, the potential energy of an electron
due to the diffusion voltage must be much larger than its average thermal
energy. Considering equation (6.13) for U D this is the case if

1< [ n;y]
ln- (6.36)

holds. At sufficiently high doping concentrations the Schottky approxima-


tion is therefore always applicable provided N A and N D are of comparable
+
size, so that the numerical factors d N A / ( N A N O ) and ~ N D I ( N -t
ANg)
548 Chapter 6 . Semiconductor junctions in thermodynamic equilibrium

of relations (6.35) are close to 1. With n, = p, = 1017 cm-' and ni =


lo1' cm-3 the value 5.7 follows for the right hand side of (6.36), i.e. the
Schottky approximation is valid under these conditions.
Some consequences of the relations derived above will now be reconsid-
ered in more detail. In regard to the width w of the space charge region
at room temperature, doping concentrations ND = N A = 1OI6 cm-', and
values of E and U D corresponding to silicon, equation (6.32) approximately
yields w = 0.4 p m . In order for the space charge region of a pn-junction
to develop fully, the p- and n-regions must have lengths which clearly ex-
ceed this value. In the case above, which is typical, they must lie in the
micrometer range. This is one of the reasons why the length scale of doping
structures in electronic devices cannot be made arbitrarily small.
Equation (6.32) for w also implies that the width of the space charge
region is smaller for heavier doping of the two semiconductors. For N D =
N A = 10'' cm-3 the space charge width w is ten times smaller than for
N D = N A = 10l6 cm-'. This is understandable because larger space charge
densities enable a narrower region to bridge the diffusion voltage. For the
same reason, the n-region contributes less to the total width of the space
charge region if it is more heavily doped. An analogous statement holds
for the p-region (see Figure 6.4). Since p(0) = 0, the energy of a junction
electron has the same value at x = 0 as an electron would have in the cor-
responding infinite n- or p-type bulk semiconductor. The total potential
difference UD is distributed between the p and n-regions according to for-
mula (6.33), i.e. the contribution from the n-region is larger for larger N A ,
and that of the p-region is larger for larger N o . This is due to the larger
width of the space charge region in the material with the lower doping.
After the electric potential p(x) is known, the chemical potential p ( z ) can
be calculated by means of relation (6.14). Also the uniform Fermi level EF
of the pn-junction, which by definition equals the electrochemical potential
p(x) - ep(x), can be determined easily. One only needs to calculate p ( z ) -
ecp(z) at one particular position x where p ( x ) is known. We take x = zn.
Then it follows from equations (6.14), (6.29), (6.31) and (6.32) that

or

The uniform Fermi level of the junction is uniquely determined by the indi-
vidual Fermi levels and doping concentrations of the n- and p-regions. The
fact that p - and n-type materials exist at a semiconductor junction despite
the common, uniform Fermi energy, is due to the bending of the conduction
and valence band edges. This bending is such that, on the left-hand side
6.2. Heterojunctions 549

Figure 6.6: Lineup of the band edges and the chemical potentials for a pn-junction
with N D = 2N.4. The chemical potential p ( z ) and the Fermi level are also shown.

of Figure 6.6, the valence band edge is close to the Fermi level so that the
material is p-type there, and on the right-hand side it is the conduction band
edge that is close to the Fermi level so that material is n-type there.

e6.2 Heterojunctions
The pn-junction is a semiconductor structure based on a single semiconduc-
tor material, except the state of the semiconductor, specifically, the value of
the chemical potential, is different on the two sides of the junction. In the
case of semiconductor heterojunctions not only are the states different but
also the materials differ - left of the nominal transition at z = 0 one has
a particular material 1, and right of it there is a different material 2. The
material inhomogeneity means that the valence and conduction band edges
E,, E,, the energy gap E,, the effective band densities of states N,, N,, the
Fermi level E F and other quantities have different values on the two sides
550 Chapter 6 . Semiconductorjunctions in thermodynamic equilibrium

Material 1
I *

Figure 6.7: Semiconductor heterojunction.

of the junction (see Figure 6.7). We indicate these differences by a material


index, i.e. we write Evl, Ev2, Eel, Ecz etc.
Heterostructures of this kind have already been encountered in section 3.7,
where they were seen to play a key role in semiconductor microstructures.
The discussion of the electronic structure of heterostructures, in particular,
their valence and conduction band lineups as well as their band discontinu-
ities, need not be repeated here. We will concentrate on the effects of free
carriers at heterojunctions. Such carriers were absent in most of the consid-
erations of section 3.7, and to the extent that they were discussed, results
were merely quoted without derivation. Now, we will present a detailed anal-
ysis of free carrier effects at heterostructures, similar to that above for the
pn-junction. As in the latter case, we initiate this analysis with discussion
of the thermodynamic equilibrium state of the free carrier system.

6.2.1 Equilibrium condition


As seen above, the thermodynamic equilibrium state of the free carrier sys-
tem of an inhomogeneously doped semiconductor is characterized by the
condition that its electrochemical potential is uniform in space. Here, we
generalize this condition to heterostructures, i.e. to semiconductor systems
having not only inhomogeneous doping but also inhomogeneous materials
composition. Again, the characterization of thermal equilibrium is focused
on the vanishing of the net electron and hole currents. However, the phe-
nomenological equations for these currents differ in the case of systems hav-
ing heterogeneous materials composition from those of systems involving a
single material only. This can be seen as follows. Assume for the moment
that the band edges Ec and Ev of the heterostructure change gradually as
functions of position coordinate 2 at the interface, rather than abruptly.
6.2. Heterojunctions 551

Then, within the framework of effective mass theory, the continuous func-
tions E c ( z )and E,(z) describe position-dependent potential energies of elec-
trons and holes, and one may expect that the gradients VE, and VE, play
the role of driving forces, which result in additional contributions to the
electron and hole currents. The same may be said for the gradients V N , ( z )
and V N v ( z ) of the effective densities of states in contributing to diffusion
current. The exact forms of these current parts are not known a priori.
They can be determined either by means of Boltzmanns equation, or one
can formulate them phenomenologically on the basis of experience as we did
in the case of the electric drift and diffusion currents in equations (5.100)
and (5.101). Using Einsteins relations (5.111) and (5.112) for, respectively,
D , and D,, as well as equations (5.82) and (5.83), the electron and hole
diffusion currents in these equations can be written as u , ( k T / e n ) V n and
- u , ( k T / e p ) V p , respectively. Since the negative gradients of N c and N ,
should play a role similar to that of the gradients of n and p , we write the
generalized phenomenological equations for the electron and hole currents
as

kT
- ecp) + -Vn - (6.39)
n

1
j, = up; [O(,T~ kT
- ecp) - - 0 p
n
+ -kT
VN,
Nw 1 . (6.40)

The concentrations n and p of electrons and holes can, as above, be expressed


in terms of the chemical potential p ( z ) , where now, in contrast to equations
(6.5), the band edges and the effective densities of states are position depen-
dent. Thus,

Using these expressions for n ( z ) and p ( z ) , the two phenomenological equa-


tions (6.39) and (6.40) may be written in the more compact form

Accordingly, the vanishing of the two current densities j, and j, is guaran-


teed by the fact that the electrochemical potential p ( z ) - ecp(z) is spatially
uniform. The constant value is again referred to as Fermi energy EF. With
this, the spatial constancy of the Fermi energy E F is seen to be a prin-
cipal feature of thermodynamic equilibrium for any carrier systems, those
of single semiconductor materials with inhomogeneous doping, and those of
5 52 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

semiconductor junctions composed of different materials. In view of this, the


chemical potential follows immediately if the electric potential is known in
thermodynamic equilibrium. The latter must be determined independently.
The problems involved in such calculations are discussed in the subsection
below.

6.2.2 Electrostatic potential. GaAs/Gal-,Al,As heterojunc-


tion as an example
Again, the determination of the electric potential is based on the Poisson
equation (5.17). Generally, the dielectric constants 1 and 2 of the two
materials entering this equation are different in the present case. The dopings
of the materials may either be of the same type or of different types. Even
if they are the same type, the dopant concentrations may differ which, of
course, may also happen if the doping types are different. As in the case
of a pn-junction, one can divide the entire structure into a space charge
region about the nominal transition point at x = 0 and two electrically
neutral bulk regions on the right and left-hand side of it. The Schottky
approximation, which in the case of the pn-junction greatly simplified the
solution of the Poisson equation, is, however, generally not applicable for
heterojunctions. The solution method must be adjusted to the particulars
of the situation, as we will see below. In any case, boundary conditions
must be applied. One such condition is obtained by replacing the potential
difference between x = 00 and x = -00 by the difference of the values
E FE ~ p(+00) and E F 3 ~ p(-00) of the chemical potential at x = +oo and
x = -00, respectively. Since the electrochemical potential p(x) - ecp(x) is
spatially uniform, we have

(6.44)

The right-hand side of this equation is, in fact, the diffusion voltage in-
troduced above in the discussion of the pn-junction. The second boundary
condition is again the vanishing of the electric field at x = +00 and x = -00,
as in the case of the pn-junction. Because one of the latter two conditions
follows from the other by means of charge neutrality, it suffices to require

dio
dx l2=--00 - 0. (6.45)

The calculation of the electrostatic potential along the general lines described
above will now be demonstrated in the particular case of a heterojunction
between GaAs (material 1) and Gal-,Al,As with x < 0.4 (material 2),
illustrated in Figure 6.8. Accordingly, we assume that the relations Ec2 >
Ecl, Ncl = Nc2, and 1 = 2 t are valid. Moreover, we suppose that both
6.2. Heterojunctions 553

Figure 6.8: Lineup of the


conduction band edges at a
heterojunction between n-type
GaAs (material 1) and n-type
GaAs/Gal-,Al,As with x <
0.42 (material 2). Part (a)
shows the situation before the
establishment of equilibrium
and (c) after equilibrium has
been established. The poten-
tial variation in space is shown
in part (b) of the figure.

0 X

materials are n-type which means that N A ( z ) = p ( z ) = 0, approximately.


With these simplifications, the Poisson equation for the heterojunction reads

(6.46)

We further assume that the donor concentrations in both materials are the
~ E F of~ the two bulk semiconductors
same, so that the Fermi energies E F and
have the same energy separations from the respective conduction band edges
E,1 and Ec2 of the two materials, whence

Considering that Ec2 > Eel, we also have E F Z > E F ~and , consequently
electrons originating in material 2 diffuse into material 1. There, they accu-
mulate close to the interface, and one thus identifies an accumulation layer.
Correspondingly, in material 2, a depletion layer is formed, and n < N o
holds so that the Schottky approximation can be applied. However, in the
accumulation layer, one has n >> N o , thus the Schottky approximation is
not valid in material 1. The character of the Poisson equation here is fun-
damentally different from that of the depletion layer - the source term on
the right-hand side depends on the potential p(z) itself through the carrier
concentration n ( z ) . In fact, from relations (6.41) and (6.44), it follows that

with no1 as equilibrium carrier concentration in the infinite bulk material 1.


At large distances from the nominal transition at z = 0, the concentration
554 Chapter 6 . Semiconductor junctions in thermodynamic equilibrium

n(x) becomes n 0 1 , as one should expect. Using (6.48) the Poisson equation
takes the form

(6.49)

which is a non-linear equation that must be solved self-consistently. This


entails finding a solution p(z) which obeys the equation with the source
term depending on cp(x) as shown on the right-hand side.
Some insight into the solution may be gained as follows: We note that
the potential energy difference -e[cp(z) - p(--oo)] of an electron in the accu-
mulation layer and at --oo cannot exceed the conduction band discontinuity
AEc Ec2 - Eel, and also will not fall significantly short of it. In any case,
it is large compared to RT. Thus, in the accumulation layer, the second
term on the right side of (6.49) representing the concentration of free car-
riers, is substantially larger than N o . In these circumstances it is possible
to make a simplification which is, in a sense, complementary to that of the
Schottky approximation, namely the neglect of the concentration of b e d
charge carriers in comparison with that of the freely mobile ones. We retain
a constant value for N D , nevertheless, in our analysis below because it does
not interfere with the determination of an exact first integral of equation
(6.49), as given by

This relation is in fact a non-linear first order differential equation. The


boundary condition (6.45) is already accounted for in it. Equation (6.50)
cannot be processed further exactly. It can be solved approximately by
iteration. This means that one first substitutes an educated guess for an
approximation to the solution p(x) into the inhomogeneous right-hand side
of equation (6.50) and then solves for (dp/dx) on the left and integrates it to
determine an improved solution cpl(x). The result for the improved solution
pl(x) is then substituted into the inhomogeneous right-hand side of (6.50)
and again on solves for ( d c p l d z ) on the left and integrates it to obtain a
further improved solution p2(x). This cycle is repeated as often as needed
to achieve a desired degree of accuracy.
Our interest here lies solely in the qualitative shape of the potential cp(x)
in the accumulation layer of material 1, so we will limit our considerations
l o an estimate. The width WA of this layer, over which most of the whole
potential change in material 1 occurs, can be easily estimated from equation
(6.50). For this purpose we neglect N D and take account of the fact that the
potential drop -e[p(x) - cp(-oo)] is smaller than, but of the same order of
6.2. Heterojunctions 555

magnitude as, the conduction band discontinuity AE,. We approximately


replace -e[cp(z) - cp(-oo)] by AE,/2 in the exponential term of equation
(6.50). On the other hand, it is clear that the derivative e(dcp/dz) has the or-
der of magnitude AEJwA. Therefore, it follows from (6.50), approximately,
that

(6.51)

where the '-1' in the rectangular bracket of (6.50) has been neglected in
comparison with exp(AE,/kT). The square root factor in equation (6.51) is
the Debye length, which already played an important role in the case of the
pn-junction. There the Debye length was small compared to the width of
the space charge region, which was a depletion region for a pn-junction. For
the accumulation region of a heterojunction, the result is just the opposite
- the width of this region is reduced relative to the Debye length by a factor

which, in general, is less than 1. With AE, = 0.25 eV and k T = 25 meV


(T = 300 K ) it is 0.8. The Debye length itself has a value of about 400 A if
no1 = 10l6 ~ m - T~=, 300 K and t = 12 are assumed. Thus W A amounts to
about 320 A in the present case.
In material 2, one has a depletion region and the Schottky approximation
is applicable. The potential profile has a parabolic shape, as in equation
(6.29) for the n-doped depletion region of a pn-junction. The potential
change extends over a width which is large compared to the Debye length,
varying much more slowly than in material 1. In Figure 6.8b1the potential is
shown schematically throughout the entire heterostructure. Below it (Figure
6.8c), we plot the effective position dependent conduction band edge E,(z) -
ecp(z). Considering equations (6.44) and (6.47), we have

E
l21 - eY4-w) Ec2 - ecp(+w), (6.52)

i.e., the band discontinuity is completely compensated by the electrostatic


potential cp(x). Since this potential is a consequence of the redistribution
of free electrons which adjust under the influence of the band discontinuity
AE,, one may also say that AE, is dielectrically screened out by these
electrons.
The spatial variation of the effective band edge E, - ecp(z) in Figure
6 . 8 ~displays an interesting property, in that the GaAs accumulation layer
is bounded on the right-hand side by the Gal-,Al,As potential barrier.
Thereby, a triangular potential well is formed. The Fermi level lies above the
bottom of this well for sufficiently heavy doping, as was assumed in Figure
6.8. The average width of the well, which was estimated above to be 360 A, is
of the same order of magnitude as the width of the quantum wells considered
5 56 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

Magnetic Field Strength (Testa)-

Figure 6.9: Quantized Hall effect of the 2-dimensional electron system of a


GaAs/Gal-, A1,As-heterojunction at T = 8 m K . (After Ebert, van Klitzing,
Probst, a n d Ploog, 1982).

in section 3.7. This means that in a single GaAs/Gal_,Al,As heterojunc-


tions one can expect similar confinement effects as in Gal-,Al,As/GaAs/
Gal-,Al,As quantum well structures. In the well region, the conduction
band splits into subbands. Since the Fermi level lies above the well bottom,
these subbands are partially occupied, and the electrons of the accumula-
tion layer form a 2-dimensional electron gas. As pointed out above, one can
observe the quantized Hall effect in such a 2-dimensional gas (see Figure
6.9).
The formation of subbands has yet another implication. Namely, that
equation (6.41) for the electron concentration n ( z ) ,which is based on the
assumption that the electrons are hosted by a 3-dimensional parabolic band,
is no longer valid. Instead, the 2-dimensional subband structure of the
triangular well has to be considered. The latter can only be determined,
however, after the potential is calculated, which, in turn, cannot be accom-
plished without knowing the subband structure. Thus, a new type of self-
consistency problem emerges, beside those already considered: the Poisson
equation must be solved self-consistently with the effective mass Schrodinger
equation (6.3).
In the above discussion of the electric potential of a GaAs/Gal-,Al,As
heterojunction, a particular assumption was made concerning the dopings
of the two materials; they were both n-type, and their donor concentrations
were identical. These assumptions are not essential, however, as one can
easily see. Even if the dopant concentrations differ, a triangular quantum
well is formed. A case of particular importance is that of modulation doping
in which only the Gal_,Al,As region is n-doped while the GaAs region
6.3. Metal-semiconductor junctions 557

is undoped. In this case the electrons in the GaAs well are not exposed
to ionized donors because there are none. Therefore the mobility of these
electrons is particularly high, it can reach several millions crn2V-ls-'. An
electronic device which exploits this effect is the High Electron Mobility
Transistor (HEMT), which was already mentioned several times in somewhat
different contexts.

6.3 Metal-semiconductor junctions

Metal-semiconductor junctions are fabricated by evaporation of the metal


and the subsequent deposition of its vapor on the colder semiconductor
surface. These junctions play an important role as electrical contacts for
semiconductor samples and devices. The dependence of their resistance on
the polarity of an applied voltage can be used for electric current rectifi-
cation purposes. In GaAs MESFET's (Metal-Semiconductor-Field-Effect-
Transistors), particular metal-semiconductor junctions (so-called Schottky
contacts) also have an active electronic function: The resistance of the GaAs
layer below the metal can be varied by means of a voltage applied to the
metallic gate-electrode.

6.3.1 Energy level diagram before establishing equilibrium


The positions of the energy levels at a metal-semiconductor junction at the
instant of its formation, i.e. before the junction had enough time to relax
to thermodynamic equilibrium, are shown in Figure 6.10 schematically. All
energies are referred to the vacuum-level Eo. In Figure 6.10a a situation is
assumed in which the highest occupied energy level of the metal, which is
its Fermi energy E F M ,lies below the Fermi level EFS of the semiconductor,
which is here assumed to be n-type. The relation E F M < EFS is typical
of metals paired with n-type semiconductors, although the reverse case can
also occur (see Figure 6.10b). For p-type semiconductors, one has essentially
the same relations. The two possible cases, in which the Fermi-level of the
p-type semiconductor either lies below that of the metal, or above it, are
shown in Figures 6 . 1 0 ~and 6.10d) respectively.
The Fermi level of the metal is given by the negative of its work function
0, so that EFM = -0. The work function is defined as the smallest energy
which must be transferred to the electron system of the metal to enable
an electron to escape. It can be obtained from photoemission experiments
in which the kinetic energies &in of electrons emitted from the metal by
the absorption of photons of a particular energy hw are measured. Since
an emitted electron has absorbed one photon, its kinetic energy is Ekin =
+
hw Einitbl where Einitbl is the energy which the electron had in the metal
558 Chapter 6 . Semiconductor junctions in thermodynamic equilibrium

I L

0 X

0 X

Figure 6.10: Positional dependence of the Fermi levels of a metal and a semicon-
ductor, as well as of the band edges of the semiconductor in a metal-semiconductor
junction at the instant of its inception, before equilibrium is established. Four dif-
ferent cases are shown: n-type semiconductor (a, b), p-type semiconductor (c, d),
E F S > E F M (a, d), and EFS < E F M (b, c).

prior to emission. There are as many different values of Einitfal as there are
occupied energy levels of the metal. Thus the kinetic energies &in span the
whole occupied part of the energy spectrum of the metal, shifted by hw. The
work function represents the lower threshold energy of this distribution.
The energy separation between the conduction band of the semiconduc-
tor and the vacuum level is referred to as the electron affinity X . It, too,
is a quantity which can be determined experimentally. One measures the
photothreshold a P h of the semiconductor (see section 3.7), and subtracts the
energy gap Eg from it, obtaining the electron affinity X from the relation
X = QPh - Eg.
If a voltage U is applied to a metal-semiconductor junction, a current will
flow through it. In Chapter 7 we will show that the magnitude of this current
essentially depends on the energy EA,(M + SC) necessary, at minimum, for
6.3. Metal-semiconductor junctions 559

an electron to be transferred from the metal into the semiconductor. In the


case of a p-type semiconductor, it will be the smallest energy E,+,(M -+ SC)
necessary for the transfer of a hole from the metal into the semiconductor
that determines the current. In both cases, one calls these energies Schottky
barriers and denotes them by Q B , so that

(6.53)

For a given semiconductor material, the sum of the two Schottky barriers,
one referring to n-type doping and another referring to p-type doping, should
be independent of the metal and equal the gap energy E , of the semicon-
ductor. The Schottky barrier determines the current density j through the
metal-n-type semiconductor junction by means of the relation

(6.54)
as will be shown in Chapter 7. Here B is a coefficient which depends only
weakly on temperature. Using equation (6.54), and measuring the current-
voltage characteristics of the metal-semiconductor junction, the Schottky
barrier @ B can be determined experimentally.
Moreover, if the level ordering at the metal-n-type semiconductor junc-
tion is that shown in Figure 6.10a, then the equation

(6.55)

relates the Schottky barrier Q B of the junction, the work function Q of the
metal and the electron affinity X of the semiconductor. In employing this
relation we have, however, jumped ahead of the thermodynamic analysis
of energy level ordering at the metal-semiconductor junction. In Figure
6.10, in fact, the energy levels are shown at the instant when the metal-
semiconductor junction has just been formed. Since the chemical potentials
of the metal and the semiconductor are different, the situation shown in
Figure 6.10 surely does not represent thermodynamic equilibrium. Electrons
or holes of the semiconductor will diffuse into the metal (Figure 6.10a,c) or
from the metal into the semiconductor (Figure 6.10b,d). Thereby a space
charge region will arise at the interface (see Figure 6.11) with its concomitant
electrostatic potential cp(z) varying in space. This potential will now be
considered in more detail.

6.3.2 Electrostatic potential


In equilibrium, the electrochemical potential EF = p ( z ) - ecp(z) is spatially
uniform. In the metal, the spatial changes of p ( z ) and p(z) are limited to a
560 Chapter 6 . Semiconductorjunctions in thermodynamic equilibrium

M S
I \ *
X

very thin boundary layer at the interface. The thickness of this layer is about
one A. This value follows from relation (6.34) if one takes the electron con-
centration of a metal to be of the order of magnitude crnV3, and replaces
the average thermal energy k T by the Fermi energy of the metal, switch-
ing over in this way from the Debye length to the Thomas-Fermi screening
length. Over a distance of an A the potential ~ ( x is) practically constant.
Deeper in the metal, because of screening by electrons in the boundary layer,
the potential is essentially constant anyway. It is therefore reasonable to ap-
proximate the potential in the metal as spatially uniform. At the nominal
transition from the metal to the semiconductor, the potential could be un-
derstood as in the case of semiconductor heterojunctions considered in
~

the previous section - to exhibit a discontinuity [cp(+O) - cp(-O)] because of


the formation of a dipole layer. At the outset, we wish to ignore this and
explore the possibility that the potential cp(;c) may be continuous at 2 = 0,
i.e. assume that
Cp(+O) - cp(-0) = 0. (6.56)
With this, it is clear that expression (6.55) for the Schottky barrier Cpg
will be valid even after thermodynamic equilibrium is established at the
junction by the formation of a space charge layer and its associated potential.
Both levels, the Fermi level of the metal and the conduction band edge, are
lifted by the same energy -ecp(O), so that their difference remains unchanged
(Figure 6.12).
Between the metal (z 5 0) and the bulk region of the semiconductor at
x = 00, a potential difference

U D = Cp(m) - Cp(0) (6.57)


evolves which corresponds to the diffusion voltage of a pn-junction. The
6.3. Metal-semiconductor junctions 56 1

E, -elp( x 1

I c
0 X

I b

0 X

Figu 2 6.12: Positional dependence of the energy levels of the metal-semiconducta


junction of Figure 6.11 after equilibrium is established. In cases (a) and (c) one has
non-ohmic contacts, and in cases (b) and (d) ohmic contacts.

spatial uniformity of the electrochemical potential implies that

If the Fermi energy of the semiconductor lies close to the conduction band
bottom, the diffusion voltage approximately equals the Schottky barrier QB.
The latter typically has values of several tenths e V , which are also typical
values that one should expect for the diffusion voltage U D .
The details of the spatial dependence of the potential p(z) depend on the
relative positions of the Fermi levels in the metal and semiconductor. For an
n-type semiconductor, if EFS > EFM holds, then electrons transfer from the
semiconductor into the metal and an electron depletion layer develops in the
semiconductor close to the interface. The diffusion voltage U D is positive,
562 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

i.e. the two bands are bent up at the interface (Figure 6.12a). One speaks of
a non-ohmic or Schottky contact in this case. If, for a n-type semiconductor,
the reverse ordering E F S < E F M holds, then electrons from the metal trans-
fer into the semiconductor, and an electron accumulation layer develops in
the semiconductor at the interface (Figure 6.12b). The diffusion voltage U D
is negative, i.e., the bands are bent down at the interface, as in the case of
the heterojunction of Figure 6.8, expect that here there is no development
of a potential well in which electrons are confined to two dimensions. This
is called an ohmic contact.
Consider next p-type semiconductors. For EFS > EFM there is a hole
accumulation layer, because electrons transfer from the semiconductor into
the metal or, what is the same, holes transfer from the metal into the semi-
conductor. The diffusion voltage U D is positive and the bands are bent
up at the interface (ohmic contact, see Figure 6.12d). For EFS < E F M ,
there is a hole depletion layer, U D is negative, and the bands are bent down
(non-ohmic contact, see Figure 6.12~).Below we first consider metal-n-type
semiconductor contacts which are non-ohmic.

Non-ohmic metal-n-type semiconductor contacts.

The potential cp(x) can be calculated in the same way as for a pn-junction. In
particular, the Schottky approximation can, again, be used, provided that
the conditions (6.35) are satisfied. We denote the width of the depletion
layer in the semiconductor by XB. The arbitrary constant of the potential
will be chosen such that cp(x) vanishes deep in the semiconductor.
cp(x) Iz=w= 0. (6.59)
The two boundary conditions for p(x) are taken to be the same as in the
case of the pn-junction. Firstly, the potential difference c p ( 0 0 ) - cp(0) must
match the diffusion voltage U D in accordance with (6.58). Secondly, the field
strength deep in the semiconductor approaches zero, i.e.

-
dcp 12=w= 0. (6.60)
dx
With these conditions the potential cp(x) and the width X B of the depletion
layer can be calculated from the Poisson equation, with the result

(6.62)
6.3. Metal-semiconductor junctions 563

With a diffusion voltage of several tenths e V , the width X B of the space


charge region takes values which correspond to those of a pn-junction for
comparable doping concentrations. For N o = 1OI6 ~ r n - ~X ,B is several
tenths of ,urn or less.
The case of ohmic metal-n-type semiconductor contacts will not be treated
here in detail. Qualitatively, the conditions in the semiconductor are similar
to those in the accumulation layer of a heterojunction considered in section
6.2. The potential energy - e v ( x ) of an electron in the semiconductor in-
creases very rapidly, i.e. over a distance which is small compared to the
Debye length, from its negative value -eUD at the interface up to a value
close to zero in the semiconductor bulk (see Figure 6.12b).
The potential profile of a metal-p-type semiconductor junction is analo-
gous to that of a metal-n-type semiconductor junction, as shown in Figures
6.12c,d.

6.3.3 Schottky barrier


The importance of the Schottky barrier @B for the current-voltage charac-
teristics of a metal-semiconductor junction was pointed out at the beginning
of this section. Under the assumption (6.56) that excludes discontinuities
of the potential at the interface, equation (6.55) is valid for @B. We refer
to it as the Mott relation since Mott was the first to propose it. In Figure
6.13, Schottky barriers are shown for a series of n-type-Si-metal junctions
as a function of the metal work function a. The electron affinity of silicon
is 4.05 e V . The straight line in Figure 6.13 corresponds to the prediction
of Motts relation (6.55). One readily sees that the linear variation of @B
as a function of @ with slope 1, predicted by this relation, is not at all ful-
filled. Rather, an approximate independence of the Schottky barrier and
the metal work function can be recognized from the figure. One may con-
clude from this that the assumption of continuity of the potential at the
metal-semiconductor interface is not valid, just as in the case of a semicon-
ductor heterojunction. Admitting a potential discontinuity to consideration,
equation (6.55) is supplanted by the the generalized relation

@B = @ -X - e[cp(+O) - p(-O)]. (6.63)


It is important to recognize that a discontinuous potential jump occurs only
if a dipole layer exists. Such a dipole layer at a semiconductor-metal in-
terface can arise as a result of interface states, i.e. electron states of the
semiconductor which are localized at the interface and have energy levels in
the semiconductor gap. The existence of such states follows from the same
considerations which led us to predict the occurrence of surface states in
section 3.6.
564 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

Figure 6.13: Schottky barriers


for various metal-n-Si junctions
versus the work function of the
metal (After Bechstedt and En-
derlein, 1988.)

Work Function @ (eV)-

Below, we will assume that interface states are also present at the metal-
semiconductor junction. We denote their energy levels with respect to the
vacuum level by EA. The concentration of the semiconductor interface state
levels will be taken to be sufficiently large to host all electrons which, in
their absence, would otherwise occupy states at higher energies. We will
show that under these circumstances, the energy EL of the interface states
is also the electrochemical potential, i.e. the energy to which the common
Fermi level of the metal-semiconductor junction adjusts in equilibrium The
proof rests on the general relation (4.57) between electron concentration,
Fermi distribution and density of states, which will be applied to the case
under consideration. For simplicity, we consider zero temperature T , but the
conclusions to be derived also hold for T > 0. The semiconductor will be
taken as n-type, and the energy levels EL of the interface states lie below the
Fermi energy of the bulk semiconductor. In reality, the interface levels do
not concentrate exactly at E i , but are spread over a narrow interval about
EL. The interface state density is therefore not an exact &-function, but a
bell curve of finite width.

Conceptually, we regard the process of establishing thermodynamic equi-


librium at the metal-semiconductor junction as taking place in two steps,
firstly, the establishment of equilibrium within the semiconductor, i.e. be-
tween the semiconductor bulk and the interface at the metal, and secondly,
the establishment of equilibrium across the interface between the semicon-
ductor and metal. In the first step, electrons transfer from the semiconductor
bulk into the interface levels. Thereby, these levels fill starting from below
up to a certain maximum energy. By definition, this maximum energy rep-
6.3. Metal-semiconductor junctions 565

resents the Fermi energy of the semiconductor. Since it lies within the bell
curve of the state density it practically equals Ei. An increase or reduction
of electron concentration in the semiconductor bulk by changing its doping
will increase or reduce the number of electrons filling the interface levels.
Because of the assumed large density of interface states, however, the posi-
tion of the Fermi level needs to change only very little to account for such
an increase or reduction. Thus, the Fermi level is fixed at the interface state
energy levels. This is called a pinning of the Fermi level.
In the second step, the establishment of equilibrium across the interface
between the metal and the semiconductor, electrons are transferred across
the interface, strictly speaking between the thin charged boundary layer in
the metal and the interface states of the semiconductor. Thereby, a dipole
layer is formed at the interface. This is accompanied by a potential jump
[cp(+O) - p(-O)] between semiconductor and metal, which in equilibrium is
sufficiently large that the Fermi level of the metal is raised or, respectively,
lowered, to match that of the semiconductor, whence

E F M - ecp(-O) = EL - ecp(+O). (6.64)


Combining relations (6.63) and (6.64), and using EFM = -@, we obtain
@ B = - E i - X . Referring the interface levels EL to the top of the valence
band rather than to the vacuum level as done so far, the redefined value of
the interface levels is E , = EL - E,. Using - X - E, = E,, these relations
yield

@B = E , - E,. (6.65)
The latter expression for the Schottky barrier was first derived by Bardeen.
In it, the Schottky barrier, in contrast to Motts expression, is seen to be
independent of the work function of the metal. Figure 6.13 shows that the
experimental data for Si are closer to Bardeens relation than to that of Mott.
This statement also holds for other group-IV semiconductors, such as Ge.
For the 111-V semiconductors, the dependence of the Schottky barrier @ B
on @ is noticeable, although smaller than that in Motts relation, and for II-
VI semiconductors it corresponds approximately to Motts linear dependence
with slope 1. These observations suggest a relation for @B which interpolates
between Motts and Bardeens formulas. Accordingly, we write

@B= S(@ + E F S )+ Po, (6.66)


where S is a slope parameter which varies between 0 and 1, @o is a constant,
and EFS is the Fermi energy of the semiconductor with respect to thevacuum
level. For n-type semiconductors, EFS approximately equals - X . The Mott
relation follows from (6.66) if one sets S = 1 and @o = 0, and the Bardeen
relation emerges if one takes S = 0 and = Eg - E,.
566 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

Figure 6.14: Dependence of the A t


slope parameter S on electroneg- c-------
0

ativity difference (After Kurtin,


McGill, and Mead, 1969.)
I I
I zns ZnC

I
0.6 )CdS

1 -
0 0.4 0.8 1.2
Electronegativity difference -
1.6 2. I

In accordance with above remarks about Schottky barriers for the various
semiconductor material classes, the value of S should increase monotonically
starting from the covalent group-IV semiconductors Si and Ge through the
moderately polar 111-V semiconductors up to the strongly polar 11-VI semi-
conductors. To examine this conjecture, S is plotted in Figure 6.14 as a
function of the electronegativity difference between the two types of atoms
paired in a given semiconductor, which represents a measure of the polar-
ity of the semiconductor material. The results verify the expected behavior
of S. The different properties of the interfaces of various semiconductors
with metals are, therefore, reflected by S in an overall way. Interface effects
are evidently more pronounced in the case of completely or preferentially
covalently bound semiconductors than in the case of more ionically bound
materials.
Experimental investigations show that Schottky barriers also depend on
the chemical and structural imperfections of a particular interface. This
means that, depending on the preparation of the interfaces, different Schott-
ky barriers may be obtained for the same metal-semiconductor junction.
Thus, for a theoretical understanding of Schottky barriers, it is particu-
larly important to study absolutely pure and structurally perfect metal-
semiconductor junctions, fabricated under UHV-conditions. From experi-
mental investigations of such junctions as, for example, with GaAs, con-
clusions have been drawn about the nature of the interface states assumed
in Bardeen's model. In the unified defect-model by Spicer for 111-V semi-
conductors, it is assumed that these states are due to point defects of the
semiconductor like vacancies or anti-site defects being induced by positioning
the metal on top of the semiconductor. In all, the microscopic understanding
of semiconductor-metal junctions still needs further investigation.
6.4. Insulator-semiconductorjunctions 567

6.4 Insulator-semiconductor junctions


Interest in insulator-semiconductor interfaces is, on the one hand, of purely
scientific origin - one wants to understand the particular macroscopic and
microscopic properties of these junctions. On the other hand, insulator-
semiconductor interfaces are of great importance for electronic semiconduc-
tor devices. In such devices, they are either used as layers for isolation or
passivation purposes, or they are involved in active electronic functions, such
as tuning the resistance of a lower-lying semiconductor layer, as occurs in the
case of the Si-MOSFET (Metal-Oxide-Field-Effect-Transistor). Without the
latter device, modern highly integrated microelectronics would not be con-
ceivable. Therefore, the insulator-semiconductor junction, particularly that
between SiOz and Si, represents one of the technological most important
solid state systems of all. Here, we will discuss the properties of insulator-
semiconductor junctions in thermodynamic equilibrium. For MOSFET op-
eration, the changes of these properties due to the application of a voltage
between the insulator and the semiconductor are important, and these will
be treated in section 7.3.

6.4.1 Thermodynamic equilibrium


Just as insulators differ from semiconductors only by the size of the energy
gap, i.e. quantitatively, the differences between semiconductor heterojunc-
tions and insulator-semiconductor junctions are mainly of a quantitative
nature. A certain qualitative difference exists in the occurrence of interface
states which is more likely for insulator-semiconductor junctions than for
semiconductor- heterojunctions. Initially, we will omit interface states from
consideration. The energy diagram of an insulator-semiconductor junction
at the instant of its inception, i.e. before establishing equilibrium, is shown
in Figure 6.15. The depicted situation, in which the energy gap of the semi-
conductor lies completely within the gap of the insulator, is not the only
possible one, but it is typical. The SiOz/Si-junction is of this type. The
relative positions of the valence bands of the insulator and semiconductor,
and the relative positions of their conduction bands, may be characterized
in the same way as was done in the case of heterojunctions. Again, the
band discontinuities cannot be determined by simply comparing the bulk
values of the band edges, since they are modified by the discontinuity of the
electrostatic potential induced by the dipole layer at the interface.
Immediately after fabrication of the junction between an insulator and
a semiconductor, the chemical potentials of the two materials have different
positions. Equilibrium adjustments involve the exchange of electrons and
holes across the interface. This leads to a space charge region and an elec-
trostatic potential p(r) which equilibrates the inhomogeneity of the chemical
568 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

Figure 6.15: Lineup of the band edges and the Fermi levels at an insulator-
semiconductor junction (a) before and (b) after equilibrium is established.

potential, such that the electrochemical potential E F = p(z)- ecp(z) is spa-


tially uniform in the final equilibrium state.
We explore the positional dependence of the electrostatic potential p(z)
below. To be specific, we assume a n-type semiconductor, and suppose that
the free carriers in the insulator are predominantly holes. Then we have a
pn-junction which differs, however, from that considered in section 5.1 be-
cause it is composed of two different materials. Another particular feature
of this pn-junction is that the acceptor concentration NAI of the insula-
tor is smaller by several orders of magnitude than the donor concentration
Nos of the semiconductor. It is natural to assume that the Fermi level
EFS of the semiconductor lies above the Fermi level E F I of the insulator.
Then, in order to establish equilibrium, electrons have to transfer from the
semiconductor into the insulator. Consequently, the semiconductor becomes
positively charged and the insulator is negatively charged.
The Poisson equation for the concomitant potential cp(z) has the general
form (6.21). The material inhomogeneity of the junction enters this equation
through the relation between the electron concentration TI(.) and the poten-
tial cp(z) as well as through the static dielectric constant which has different
values 1 and c in, respectively, the insulator and the semiconductor. If n ( z )
does not appear explicitly, as occurs within the Schottky approximation,
the material inhomogeneity simply results in the replacement of ( N A / E )by
( N A / E I ) and
, ( N o / ) remains ( N o / ) .
Below, we employ the Schottky approximation and later verify its justi-
fication. The normalization and boundary conditions are again chosen ac-
cording to relations (6.9) and, respectively, (6.22) and (6.23). The continuity
of electric field strength - ( d p / d x ) at x = 0 is replaced by the continuity of
6.4. Insulator-semiconductorjunctions 569

the electric displacement - e ( z ) ( d ~ / d z )at this position. The solution of the


Poisson equation is then given by the previously derived expression (6.29).
Relations (6.35) and (6.36), which determine the validity limits of this solu-
tion also retain their applicability. In the case considered here, N M << N D S
holds. Then, as equation (6.31) shows, the width x p of ~ the space charge
region in the insulator is considerably larger than the width 2,s of the cor-
responding region in the semiconductor. The space charge region in the
insulator can become so large that it exceeds the sample size. If this oc-
curs, the jump discontinuity of the chemical potential cannot be completely
screened out by the electrical potential in this circumstance. Hence, the
model of a junction of semi-infinite media, which underlies our calculations,
is no longer valid. The general conclusions which we derive here are not
affected by this fact. From formula (6.33) one finds that practically the en-
tire diffusion voltage drops off within the insulator. Thus the electric field
strength in the semiconductor is well approximated by zero. As far as the
common Fermi level E F is concerned, equation (6.38) shows that it adjusts
approximately to the Fermi level of the semiconductor, whence

EF = E F S . (6.67)

The application of the Schottky approximation still has to be justified. This


may be done by means of relations (6.35) which yield the following results.
For the p-region, i.e. the insulator, the Schottky approximation holds if
f i << m, which we can assume. For the n-region, i.e. the semicon-
ductor, the Schottky approximation can indeed become problematic because
+
the factor J N A / ( N A N o ) is small; however, it yields qualitatively correct
results in this case too.
Summing up and generalizing the above findings on the equilibrium state
of a insulator-semiconductor junction one can say that a relatively small
number of free carriers, transferred from the semiconductor to the insula-
tor, is sufficient to adjust the Fermi level of the insulator to that of the
semiconductor. This number is so small, viewed from the point of view of
the semiconductor, that the state of the semiconductor is almost completely
unaffected by the insulator. Moreover, because of the small space charge
density and the practically vanishing electric field in the semiconductor, the
field strength must also be small in the insulator - recall that the dielec-
tric displacement penetrates continuously into the insulator. Below, we will
neglect the electric field in the insulator completely.
In the above treatment of the insulator-semiconductor junction, interface
states were omitted from consideration. In reality, such states may exist in
the semiconductor. Their effect on the electrostatic potential of the junction
will be explored now.
5 70 Chapter 6. Semiconductor junctions in thermodynamic equilibrium

6.4.2 Influence of interface states


As in the case of the metal-semiconductor junction above, the interface states
cause the chemical potential of the semiconductor to take a different posi-
tion at the insulator-semiconductor interface than it has in the bulk. We
denote these positions by, respectively, Eg and Ek. In the case of an n-type
semiconductor and acceptor-like interface states which can host electrons, a
partially compensated semiconductor layer will arise at the interface. There,
electrons are captured by the acceptor levels, so that in the semiconductor a
negative interface charge density -ens (per unit area) is formed. This comes
about at the expense of the semiconductor bulk, where a positively charged
depletion region arises below the interface. The width x n of this depletion
region is determined by the conservation of electron number, which may be
expressed as

ns = NDSX,. (6.68)
According to electrostatic laws, the interface charge density results in a
jump discontinuity [F(+O) - F(-O)] of the electric field strength given by
-(e/tto)n. Since the field in the insulator vanishes, a non-zero field strength
F S arises on the semiconductor side of the interface, as given by

(6.69)

One can also understand this interface field as being due to the positive space
charge region in the semiconductor below the interface, which in Schottky
approximation gives rise to a field strength F ( x ) that grows linearly towards
the interface, and a potential p ( x ) which decays quadratically. The value of
the potential at the interface,

d o ) = us, (6.70)
must have just the right magnitude so that the electrochemical potential is
uniform, i.e. that the difference [Ek - E k ] between the chemical potentials
at the interface and in the bulk of the semiconductor is equilibrated. Hence,

- e U S = [ E k - E$] (6.71)
must hold. Solving the Poisson equation in Schottky approximation subject
to the boundary conditions that the potential and the field strength vanish
deep in the semiconductor at x = m, and that the potential at the interface
is given by (6.71), we obtain

(6.72)
6.4. Insulat or-semiconductor junctions 571

with

x,= /s.
eND (6.73)

Combining equations (6.68) and (6.73) yields a relation between the interface
charge density ns and the interface potential U s , which reads

(6.74)

For the interface field F S , it follows that

(6.75)

We use the latter two relations to estimate the orders of magnitude of n s


and F s ,with the resulting relations

ns = 1.05 x 103cm-2[(d/~/cm-3)(1U s I /V)]1/2, (6.76)

1 F S I = 18.97 x U s I /V)]1/2.
10-4(V/cm)[(~-1Ng/cm-3)(1 (6.77)
An estimate of about lev for e I U s I follows from (6.71) if we assume the
bulk Fermi level to be close the conduction band bottom, and the surface
Fermi level to be deep in the energy gap. With N D = cm-3 and
6 = 10 we find ns = 3 x 10l1 cmA2 and F S = 6 x lo4 V/cm. The width

zn of the depletion region is 0.3 pm. Thus, it is to be expected that the


fields and interface charge densities are of considerable size at an insulator-
semiconductor interface.
Our assumption that the Fermi level at the interface lies deep in the
energy gap is justified only if the sheet concentration of interface states has
at least the same order of magnitude as the calculated electron concentration
ns, i.e. the order of magnitude lo1' cm-'. This may be valid in some cases,
but in others it is definitely not. An example of the latter case is the interface
between SiOa and Si, which is largely free of interface states. As we will see
in Chapter 7, this is decisive for the operation of the MOSFETs, - indeed,
it is downright necessary to make this device possible. A large interface
charge density means, as we know from section 6.3, that the Fermi level
is pinned at the energy of these states, while MOSFET operation depends
upon having the capability to tune the Fermi level by means of an external
voltage. The interface of GaAs with its own oxide or with foreign insulators,
behaves otherwise: The interface state density is large, hence MOSFETs
made of GaAs are difficult to achieve. In the case of GaAs one must turn to
the MESFET to fabricate a field effect transistor.
572 Chapter 6. Semiconductorjunctions in thermodynamic equilibrium

Figure 6.16: Lineup of the band


edges of an insulator-semiconductor
junction in equilibrium in the pres-
ence of interface states.

6.4.3 Semiconductor surfaces


Everything discussed above pertaining to insulator-semiconductor junctions
also holds analogously for free semiconductor surfaces. In this regard, the
insulator is replaced, so to speak, by vacuum, which may be thought of as
a special insulator with an infinitely large energy gap and (relative) static
dielectric constant 6 of unity. The positional dependence of the energy levels
after equilibrium is established is essentially given by Figure 6.16, except
that the potential curve on the insulator side is now exactly zero, since no
free carriers exist in vacuum.
Instead of interface states, surface states can occur, as is known from
section 3.6. These have the same effect that interface states have in the
case of the insulator-semiconductor junction. They result in the formation
of a surface charge density n s , which is accompanied by a surface field F S
and a surface potential U s . For these quantities, equations (6.71), (6.74)
and (6.75) hold without change. Therefore, at free semiconductor surfaces,
sheet charges and electric fields of considerable size occur, just as in the
case of an insulator-semiconductor junction. These are utilized in surface
photoeffects, wherein optically excited free carriers screen the surface field,
partially or completely, depending on the intensity of the absorbed light.
Photoreflectance is a kind of electroreflectance (see section 3.8) in which
this effect is used to modulate the surface electric field of a semiconductor
by means of light.
573

Chapter 7

Semiconductor junctions
under non-equilibrium
conditions

The treatment of semiconductor junctions above was limited to conditions


of thermodynamic equilibrium. In this chapter, non-equilibrium states of
these junctions will be explored. Such states occur when the junctions are
subjected to various external perturbations, which can include the applica-
tion of an electric voltage, irradiation with light, exposure to pressure, or the
introduction of a temperature gradient. Here, we limit our considerations to
the application of a voltage and irradiation with light, because they are by
far the most important perturb at ions generating non-equili br ium conditions
in semiconductor junctions. In sections 7.1 and 7.2 we treat the pn-junction
subject to an external voltage. The interaction with light is considered in
section 7.2. The influence of a voltage on a metal-semiconductor junction
forms the subject of section 7.3, and in section 7.4 we examine voltage effects
on an insulator-semiconductor junction.
In the introduction to Chapter 6, we pointed out that perturbation of the
equilibrium state of a semiconductor junction provides the basic mechanism
of operation for most semiconductor devices. Examination of the effects of
an applied voltage and interaction with light for such junctions therefore con-
stitutes an exploration of the physical principles of a number of such devices.
In section 7.1, the devices involved are rectifying semiconductor diodes, the
bipolar transistor and tunnel diode; in section 7.2 the photodiode, the so-
lar cell and the semiconductor injection laser. Section 7.3 deals with the
metal-semiconductor rectifier, and section 7.4 with the field effect transis-
tor. In this preview of the content of the present chapter, it is worthwhile
to note that we have omitted perturbation of the equilibrium of semicon-
ductor heterojunctions. There are two reasons for this omission. The first
5 74 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

concerns conventional semiconductor devices. Although heterojunctions are


used in many of them, in none do they play a decisive role for the operation
of the device. They mainly just improve its operation as, for example, in
the case of an injection laser using a quantum well instead of a homogeneous
material region as active layer. The second reason concerns devices whose
principle of operation actually relies on heterojunctions. As an example, we
recall the High Electron Mobility Transistor (HEMT) mentioned in Chap-
ter 6 which utilizes the %dimensional electron gas of the quantum well at a
GaAs/(Al,Ga)As interface. While the (A1,Ga)As alloy region is n-doped to
provide the electrons for the 2-dimensional gas, the GaAs well region itself
is undoped to avoid carrier scattering by ionized donors which would lower
the mobility. The detailed treatment of such quantum devices which, with a
few exceptions (the HEMT among them), are currently subject to research
would exceed the framework of this book as an introduction to the principles
of the semiconductor physics and devices. Therefore, we omit them here.
Readers interested in quantum devices will find an overview in a recently
published monograph (Kelly, 1995).
Restricting external perturbations to the application of a voltage and
irradiation by light means that devices which are sensitive to physical influ-
ences other than electricity and light, like pressure, temperature, chemical
concentrations or magnetic fields, are ruled out. Such devices can convert
non-electric signals into electrical signals. They are called semiconductor
sensors. One of them, the photodetector, which involves the conversion of
light, will be treated in section 7.2.

7.1 pn-junction in an external voltage


The application of an external voltage ( s e e Figure 7.1) causes the free charge
carriers of the pn-junction, which originally had been in global thermody-
namic equilibrium, to be driven into a global non-equilibrium state. To
start, it is important to clarify how this state can be adequately described.
It turns out that the understanding of non-equilibrium states used above,
which rests on the concept of local equilibrium and a spatially varying chem-
ical potential p ( x ) , must be modified in an important respect if applied to
biased pn-junctions.
Local equilibrium means, in particular, that a local chemical potential
p ( x ) exists for the entire system of free carriers, including both electrons
and holes. That this can no longer hold when a voltage is applied to the pn-
junction is easily understood. Indeed, on that side of the junction where the
potential energy of electrons is raised by the applied voltage, the electron
concentration will decrease compared to the unbiased case. On the same
side, the potential energy of the holes is necessarily lowered and their con-
7.1. pn-junction in an external voltage 575

Figure 7.1: pn-junction under an


applied voltage.

centration is raised in the presence of the voltage. In view of the relations


( 6 . 5 ) between p ( 2 ) and the carrier concentrations n ( z ) , p ( z ) , this means
that, on the above considered side of the pn-junction, the chemical potential
of the electrons must be lower, and that of the holes must be higher than
for the unbiased junction. In any case, the chemical potential p n of elec-
trons differs from the chemical potential pP of holes provided - and this is
an essential restriction - these potentials exist at all.
In order to clarify whether they do exist or not, we recall that the chem-
ical potential is defined as a parameter of an equilibrium distribution func-
tion. By assuming different chemical potentials for electrons and holes, the
energy distributions of these carriers within their respective bands are sup-
posed to be in equilibrium. Only the distribution of carriers between the
bands does not correspond to an equilibrium state if the chemical poten-
tials are different. This assumption is justified if the processes which tend
to establish equilibrium proceed much faster among electrons in the same
band than among electrons in different bands. In the first case one has en-
ergy relaxation of the carriers in their respective bands. The characteristic
energy relaxation time is typically of the order of magnitude lo-'' s. In
the second case one has capture or recombination with a characteristic time
typically of the order of magnitude lo-' s, i.e. 100 times larger than the
energy relaxation time. In such circumstances it is reasonable to assume
that the two types of carriers are in equilibrium with respect to the energy
distribution within their respective bands, even if no equilibrium exists with
respect to their distribution between these bands. The assumption of sepa-
rate chemical potentials for electrons and holes is also justified under these
conditions.
At a pn-junction, both chemical potentials are functions p n ( 2 ) and pP(;c)
of position. These functions must be determined separately in conjunction
with the electric potential cp(z). We first solve for cp(z), and later for the two
chemical potentials. The concentrations n ( z ) of electrons and p(z) of holes
needed in these calculations are, by definition of the separate local chemical
potentials, given by the relations
5 76 Chapter 7. Semiconductor junctions under non-equilibrium conditions

7.1.1 Electrostatic potential profile


Formally, the presence of an external voltage U manifests itself in a change
of the potential drop across the pn-junction, which in equilibrium was the
diffusion voltage U g , to the new value

Here, the external voltage U , as usual, is counted as positive if it leads to a


decrease of the potential energy of a positive charge carrier at z = +m as
compared to x = -m. That the diffusion voltage does not contribute to the
voltage balance in a closed circuit containing the pn-junction was already
explained in section 6.1.
The calculation of the electrostatic potential p(z) at an unbiased p n -
junction was based upon the Poisson equation (5.19) and the Schottky ap-
proximation, which was used to eliminate the free charge carrier concen-
trations n ( z ) and p ( z ) from this equation. If one also assumes complete
ionization of the donors and acceptors in the space charge region, as done
in section 6.1, then the chemical potential does not explicitly appear in
the Poisson equation. It is only through the boundary conditions, strictly
speaking through the diffusion voltage U g , that the electrostatic potential
function depends on chemical potential in the p- and n-regions. Whether
the Schottky approximation retains validity or not in the presence of an ex-
ternal voltage must indeed be examined. If one repeats the reasoning used
to justify this approximation in section 6.1 for the unbiased case, then it is
readily seen that there are no changes, except that the diffusion voltage U D
has to be replaced by U D - U . From this we may conclude that the Schottky
approximation in the presence of an external voltage is applicable provided

holds and the doping concentrations N A and N D are of comparable size. For
positive U equation (7.4) is not always satisfied. It becomes critical when
U approaches the diffusion voltage, and is definitely invalid when U > UD.
For negative external voltages, condition (7.4) may in fact be valid even if
it does not hold in the absence of an external voltage. If it is valid in the
unbiased case, then it is even better satisfied in the presence of a negative
voltage. In the following we will assume that condition (7.4) is fulfilled and
that the doping concentrations N A , N D are comparable. Then the Schottky
7.1. pn-junction in an external voltage 577

junction on applied voltage U , -$,,


measured in units of U D . Sym-
metric doping with N A = N D = 1.0 -
10l6 ~ r n -is~ assumed. The
static dielectric constant is that 95 -
of Si (E = 12.1).

approximation is applicable, as are the concepts of the depletion and bulk


regions which were introduced in the context of an unbiased pn-junction. In
conjunction with this, all the results of section 6.1 remain valid in regard to
the potential function cp(z), except that U D must be replaced everywhere
by U D - U . For the potential values at the boundaries of the space charge
region, we have

and the width w of the space charge region is given by

In the presence of a negative voltage of 100 V , doping concentration N D =


N A = 1OI6 ~ r n -(using
~ silicon values for E and U D ) the space charge width
w is about 4 p a . For U = U D , w formally takes the value zero, i.e., no
depletion region exists at all. This signifies that the Schottky approximation
breaks down (see Figure 7.2).
The calculation of the two chemical potentials p"(z) and p p ( x ) for elec-
trons and holes, respectively, is more complicated than that of the electric
potential. It requires simplifications which are only understandable if one
already has a qualitative understanding of the nature of current transport
through the pn-junction. We will now address this point.

7.1.2 Mechanism of current transport through a pn-junction


That an applied voltage causes current flow through a pn-junction, is obvi-
ous. Less evident, however, is the manner in which it happens. Far from the
nominal transition at z = 0, one may assume that current transport does not
5 78 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

a1 u-0 Cl u>o
0 0 0 0 . .

0 0 0 0 0 0 0 0

0 0 0 0 0 . 0 0

0 0 0 0 0 0

+ l o o o 0 /. . . . ( ~
0 0 0 9 0 . 0 4 0 0 0-0 .to 0 .
c
+ ~ o o o o ( . . . .
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

0 0 0 -0 0 0-0 .to 0 o c

Figure 7.3: Mechanism of current flow through a pn-junction, a) pn-junction with-


out voltage, b) Injection of minority carriers by applying a (positive) voltage, c)
Recombination of the injected minority carriers with available majority carriers,
d) Refilling of the emptied band states by majority carriers from the bulk region.
There a majority charge carrier current flows.

differ from that in an infinite semiconductor sample, i.e. the current in the
p-region will be carried mainly by holes, and that in the n-region mainly by
electrons. In this matter, one has the peculiar situation that a hole current
in the left part of the junction passes over into an electron current in the
right part of the junction. The mechanisms which can realize such a tran-
sition are the recombination and generation of electron-hole pairs, whence
we may conclude that the latter processes should play an important role in
current flow through the pn-junction.

In our further considerations we assume CJ > 0. In order for recombina-


tion to occur, non-equilibrium carriers must be available. These appear in
the p-region in consequence of the fact that electrons move over from the n-
region to the p-region because their potential energy in the n-region is lifted
under the applied voltage by the amount eU. Therefore they overcome the
potential barrier between the n- and p-regions more easily. This transfer
of carriers across a potential barrier is called injection In the p-region the
electrons are minority charge carriers. Therefore we have an injection of mi-
nority charge carriers into thep-region. An analogous process takes place on
the n-side of the junction. There, holes from the p-region move over into the
n-region, so that one has an injection of minority holes. These relationships
are illustrated in Figure 7.3. The injected minority carriers (Figure 7.3b)
recombine with the already present majority carriers (Figure 7 . 3 ~ ) . The
7.1. pn-junction in an external voltage 5 79

states in the bands which become unoccupied in this way will be filled by
majority carriers from the bulk regions of the junction (Figure 7.3d), where
majority carrier currents therefore flow. The magnitudes of these currents
are determined by the speeds of the injection and recombination processes.
In the stationary state both speeds must be equal - all carriers which are
injected must recombine, and all carriers which recombine must have been
previously injected.
If one reverses the sign of the voltage, corresponding to U < 0, then one
has extraction of minority charge carriers instead of injection, and generation
instead of recombination. Since extraction and generation processes consume
energy, in contrast to injection and recombination wherein energy is released,
one must expect the current through the pn-junction to be much smaller for
U < 0 than for U > 0. This is in fact the case, as we will formally prove
below.
The above mechanism for current transport through a pn-junction will
now be formulated quantitatively. The total current density j ( x ) consists,
according to formula (5.85)) of the electron current density jn(x) and the
hole current density jp(x). For the two current constituents the continuity
equations (5.20) and (5.21) hold. In the present case, the generation term
is zero, and the annihilation term is determined by recombination. In the
stationary state one obtains

djp
- = -eR(x).
dx
where R ( z ) represents the recombination rate given by equation (5.61).
Adding (7.7) and (7.8), it follows that the total current density j ( x ) is free
of sources and must be spatially constant, so that

j(x) = j n ( 2 ) + jp(x) zz j = const. (7.9)


We consider a particular position X I in the p-region and another particular
position x2 in the n-region. The fact that the total current density j is
const ant yields

(7.10)
Integration of equation (7.7) provides the result

(7.11)

Combining (7.10) and (7.11)) we have


580 Chapter 7. Semiconductor junctions under non-equilibrium conditions

(7.12)

Setting
(a) 21 = -oo,z2 = +m, we find

(7.13)

On the other hand, considering

(b) z1 = x p , z 2 = zn, it follows that

j =j n ( 4 +jp(zn) +e /
XP
X,
d 4 z ) . (7.14)

We will later see that, under certain conditions, the two minority charge
carrier contributions jn(-oo) and jp(+oo) in (7.13) can be approximately
neglected. The same holds for the recombination contribution in (7.14).
Consequently, equation (7.13) means that the total current, in its essence,
represents a recombination current. Alternatively, in formula (7.14) the total
current is seen to be the sum of the minority carrier currents at the two space
charge boundaries. There, they are determined by the injection of minority
carriers. The total current represents, therefore, an injection current. The
two interpretations are equivalent, but they emphasize different aspects of
the total current.
With these considerations concerning the mechanism of the current trans-
port, we are sufficiently prepared to calculate the spatial profiles of the chem-
ical potentials of the two types of carriers.

7.1.3 Chemical potential profiles for electrons and holes


We first calculate the charge carrier concentrations n ( z ) and p ( z ) . Once
they are known, the chemical potentials follow immediately from relations
(7.1) and (7.2). Since the current is due to recombination of injected non-
equilibrium minority carriers in the bulk regions, we will restrict our con-
siderations to these carrier concentrations in particular. The space charge
region will therefore be omitted initially. Moreover, we will use the fact that
the relative change of the majority carrier concentrations caused by injection
is substantially smaller than the corresponding relative change of the minor-
ity carrier concentrations. For the majority carrier concentrations in the
bulk regions, the values without injection, i.e. without an applied voltage,
can be used. This means that
7.1. pn-junction in an external voltage 581

n(x) = n, for 2 > x,, (7.15)

p(z) = p p for x < xp. (7.16)


The minority carrier concentrations in the bulk regions are calculated from
the combined diffusion-recombination equations (5.81) and (5.87). To solve
these equations uniquely, boundary conditions are required, in the case of
electrons at -m and xp,and in the case of holes at x n and +oo. For n(-m)
and p(00) the equilibrium values

apply. The concentrations at the boundaries of the space charge region


may be taken from relations (6.15) for n ( z ) and p ( z ) . Of course, the latter
expressions were written down for equilibrium conditions, but they provide
approximately correct values for n ( z p ) and p(x,) even when a voltage is
applied. The only change to be made is the replacement of [cp(zn)- ( p ( z p ) ]
by U D - U instead of by UD. It follows that

-e(UD-u)/kT = eU/kT
n(q,) = n,e npe , (7.18)

(7.19)
For U = 0, the values n ( x p ) and p(;cn) are the minority carrier concen-
trations np and p , in equilibrium. If U > 0, the factor multiplying n p in
(7.18), and the factor multiplying p , in (7.19) is larger than 1, i.e., the
minority carrier concentrations exceed the values they would have in the
absence of an external voltage. This is the formal expression of the injec-
tion of minority charge carriers. How effective injection is can be recognized
through the following estimate. With U = 0.25 V and T = 300 K we obtain
exp(eU/kT) M elo M 2 x lo4. This is to say that the small voltage of 0.25 V
suffices to increase the minority carrier concentrations by more than 10000.
The steady state solutions of the diffusion-annihilation equations (5.90)
for n(x) and (5.97) for p ( z ) , under the respective boundary conditions (7.15),
(7.16) and (7.17), are given by

P(X) =P, + [~(z,) - P,I~-(x-x*)lLp , > ", (7.21)


where L , and L, are the diffusion lengths of, respectively, electrons and
holes. For positive external voltages one has n ( x p ) > n, and p(x,) > p,.
582 Chapter 7. Semiconductor junctions under non-equilibrium conditions

Figure 7.4: Lineup of the quasi Fermi levels of a pn-junction under an applied
voltage, (a) flow direction (U > 0 ) , (b) blocking direction (U < 0)). Dashed curves
correspond to the unbiased pn-junction. In the space charge region interpolated
values are used. The decay of the quasi Fermi levels in the two bulk regions is
drawn greatly exaggerated.

This means that in the two bulk regions, both minority carrier concentrations
are larger than their respective equilibrium values np and p,. The chemical
potential p n ( z ) of the electrons in the pregion, is therefore shifted to higher
energies with respect to its equilibrium value, and that of the holes in the
n-region is shifted to lower values. The same holds for the pertinent non-
equilibrium electrochemical potentials

E F ( ~ =) p n ( x ) - ecp(x), E;(X) = p P ( z )- ecpG), (7.22)


which are also referred to as quasi Fermi levels. In Figure 7.4a the spatial
variation of the two quasi Fermi levels is shown schematically, together with
the valence and conduction band profiles.
For a negative voltage U < 0 one has n ( z p ) < np and p ( z n ) < p,.
The chemical potential of electrons in the p-region therefore lies below the
equilibrium value, and that of the holes in the n-region lies above it. The
same is true again for the quasi Fermi levels (see Figure 7.4b). The elevation
or depression of these levels in the bulk regions is effective up to a distance
from the depletion region which roughly equals the diffusion length of the
pertinent minority carrier. In the depletion region between the two bulk
regions, we cannot make such statements, or any others, because the above
consideration excludes this region. Fortunately, the diffusion lengths are,
as a rule, an order of magnitude larger than the width of the space charge
7.1. pn-junction in an external voltage 583

region (see section 5.4), so that this lack of knowledge is not important. The
space charge regions function solely as potential barriers over which non-
equilibrium carriers are injected. The spatial expansion of the barriers can
be neglected in a first approximation.

7.1.4 Dependence of current density on voltage


We can now carry out the calculation of the total current density through
a pn-junction without difficulty, starting from the relations derived for the
recombination current density (expression 7.13, method a) and for the injec-
tion current density (expression 7.14, method b). Both methods must lead to
the same result. Formally, it would therefore suffice to consider only one of
them. We will study both to demonstrate explicitly that the recombination
current (7.13) and the injection current (7.14) are in fact identical.
Method a
From the qualitative discussion of current flow through the pn-junction in
subsection 7.1.2, we know that the currents at -co and +co do not differ
from the currents in an infinite p or n-type semiconductor. This means,
in particular, that the minority carrier currents at -00 and +co are neg-
ligibly small. In regard to the remaining integral in (7.13), recombination
processes in the bulk regions to the left and right of the depletion region
contribute significantly only up to depths which roughly equal the diffusion
lengths of the minority carriers - only there do the concentrations of these
carriers differ substantially from their equilibrium values, so it is only there
that recombination occurs. The diffusion lengths are, as has already been
mentioned, generally much larger than the width of the depletion region.
Taking advantage of this magnitude relation, we neglect the contribution of
the depletion region to the integral in equation (7.13), whence we obtain,
approximately,

(7.23)

The recombination rate R is given by relations derived above in section 5.2.


According to formula (5.64), in the presence of a small excess concentration
A n ( x ) of electrons (here, of electrons in the p-region), we have

(7.24)

and according to formula (5.66), in the presence of a small excess concentra-


tion A p ( z ) of holes (here, such in the n-region), the corresponding relation
is
584 Chapter 7. Semiconductor junctions under non-equilibrium conditions

(7.25)

with rn and rp given by (5.65) and (5.67) as minority carrier lifetimes of


electrons and holes, respectively. For An(.) and A p ( x ) we employ relations
(7.20) and (7.21) in the form

~ p ( x=) b ( x n ) - pnle-(z-xJ'Lp, xn < 2. (7.27)


Substituting these formulas in equations (7.24), (7.25) and (7.23) and using
relations (7.18) for n ( x p ) and (7.19) for p ( x n ) , we obtain

j = js(e
eUlkT - 1) (7.28)
with
(7.29)

The same result is obtained if one proceeds in accordance with method b,


which will be verified below.
Method b
In equation (7.14) for j we again neglect the recombination integral over the
depletion region. The minority carrier current densities j n ( z p )and j p ( x n )at
the boundaries of the depletion region follow from the general phenomeno-
logical equations (5.100) and (5.101), wherein the electric field strength has
to be set zero since the electric potential is constant in the bulk regions.
The charge carrier concentrations n ( x ) and p ( x ) are taken from expressions
(7.20) and (7.21), yielding

(7.31)

Adding these equations and applying relations (7.18) and (7.19), the expres-
sion (7.28) follows for j with

(7.32)

The two expressions (7.32) and (7.29) for j , are identical, since Ln =
(see equation 5.95) and L, = fi (see equation 5.99).
7.1. pn-junction in an external voltage 585

Figure 7.5: Current-voltage char-


acteristic of a pn-diode made of Ge.
Note the different voltage scales for
forward and reverse biasing. For ex-
tremely large negative voltages, the
diode undergoes electrical break-
down. (After Seeger, 1973.)

With this observation, the task of calculating the current through a pn-
junction under an applied voltage is completely solved, within the framework
of the conditions and approximations set forth above. We will now discuss
the results. The current-voltage characteristic (7.28) is extremely non-linear.
It exhibits the expected asymmetry with respect to a change of the sign of
voltage U . For a positive U of a few tenths of a Volt, j is several orders of
magnitude larger than j,, while for a negative U of same absolute value, j
approaches - j , (see Figure 7.5). The current density j,, which cannot be
surpassed at even larger negative voltages, is called the saturation current
density. To estimate the size of this current density in the case of Si, we
assume typical values for L,,p of 10 pm and for 7n,pof lo-' s. For the
minority carrier concentrations, we obtain, from np = n?/NA and p, =
n!/ND with ni = lo1' cm-3 and N A = N D = 1OI6 cmV3, the values np =
p, = lo4 c ~ L - Using
~. e = 1.6 x 10-lgA s, it follows that j, 10-l' A/cmz.
The saturation current density is therefore extremely small. Thus, for U < 0
practically no current flows, the pn-junction blocks the current flow. One
says that it is reverse biased or biased in blocking direction, The biasing
U > 0 refers to the forward bias or Bow direction because the current density
in this direction is orders of magnitude larger than j, as we have seen above.
The pn-junction operates as an electrical rectifier. It is called a rectifyingpn-
diode in this context. In Figure 7.5 a measured current-voltage characteristic
of a pn-diode made of Ge is shown.

7.1.5 Bipolar transistor


The pn-junction has important application in the bipolar transistor. This
device consists either of two n-regions, which are separated by a p-region, or
of two p-regions separated by an n-region (see Figure 7.6). In the first case
one speaks of an npn-transistor, and in the second case of a pnp-transistor.
586 Chapter 7. Semiconductor junctions under non-equilibrium conditions

Figure 7.6: Bipolar npn- and pnp-transistor.

m Emitter Base Collector

Figure 7.7: npn-transistor in the common-emitter configuration (left-hand side).


Illustration of the current flow (right-hand side).

In the following considerations we confine our attention to the npn-case.


In Figure 7.7 the npn-transistor is shown in one of the possible switching
modes, called common-emitter configuration (for reasons explained below).
The left n-region is connected both with the p-region in the middle as well
as with the n-region on the right. The voltage source of the np-circuit puts
the left n-region at a potential P E , and the p-region at a potential p ~ The
.
voltage source in the npn-circuit puts the potential of the right n-region at
(pc.We consider the case in which

PE PB (PCI (7.33)
and, accordingly,

(7.34)

holds. In this case the left pn-junction is biased in the flow direction, and the
right in blocking direction (see Figure 7.8). From the left n-region, electrons
are injected into the p-region in the middle, while the right n-region extracts
electrons from the p-region. One therefore calls the left n-region the emitter,
and the right the collector. The p-region in the middle is called the base.
Accordingly, the np-circuit will be referred to as the emitter-base circuit or,
in short, the base circuit and the npn-circuit as the emitter-collector circuit,
7.1. pn-junction in an external voltage 587

or in short, the collector circuit. Let the current in the base circuit be ig,
and that in the collector circuit ic. Then the current i E through the emitter
follows from Kirchhoffs current branching theorem as the sum of the two
currents,

i~ = iB + ic, (7.35)
One can also say that the emitter current splits in two partial currents, one
flowing through the base, and one flowing through the collector (see Figure
7.7 on the right).
Our goal is the calculation of the three currents i ~ iB, , ic. In this matter,
we can employ the results obtained above for the current flow through an
individual pn-junction, with the valence and conduction band edges of the
npn-transistor lined up as shown in Figure 7.8. The pn-junction on the
emitter side is subject to the flow voltage U g , and the pn-junction on the
collector side is subject to the blocking voltage U c - UB. The emitter current
i~ is the current flowing through the emitter-side pn-junction. As such it is
given by expression (7.12), multiplied by the emitter area A . Accordingly,
it consists of the injection current j n ( z p ) Aof the minority electrons into the
588 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

base, the injection current jp(x,)A of the minority holes into the emitter,
and the recombination current in the depletion region. Neglecting the latter
contribution as before, and using relations (7.30) and (7.31), we have

Here the signs are opposite to those of relations (7.30) and (7.31) above, due
to the fact that the emitter-base junction has thep-region on the right-hand
side and the n-region on the left, whereas it was the opposite above. If
both regions were expanded infinitely, as we always assumed above in the
treatment of the pn-junction, then the minority charge carrier concentra-
tions n(x) and p(x) in (7.37) could be replaced by the previously derived
expressions (7.20) and (7.21). This procedure would result in expressions
for j,(xp) and j p ( z n )which are just the negatives of those in relations (7.30)
and (7.31). In the transistor, however, only the emitter can still be consid-
ered to be infinitely extended, while the width b of the base must be treated
as finite because it is not large in comparison with the diffusion length L, of
the minority charge carriers. Thus, only the injection current density jp(x,)
of holes may be taken from the previously derived expression (7.31). Ad-
justing this expression to the relationships at the emitter-base junction, we
find along with (7.27),

(7.38)

The injection current density jn(zp)of electrons, however, must be calculated


anew. This can be done on the basis of the following considerations. To
start, it is clear that, because of the finite width of the base, only part of the
injected minority electrons recombine in the p-region, while the remainder
diffuse through this region and reach the depletion region at the collector side
of the pn-junction. The negative electric field there pulls the electrons into
the bulk region of the collector, from which they are sucked up by the applied
positive voltage. The equilibrium value np of the electron concentration is
therefore reached not only at x = 00, but it is already realized at x = b. This
means that the boundary condition An(x = b) = 0 must now be imposed.
The solution of the diffusion-recombination equation (5.91), which accounts
for this new boundary condition, results in

(7.39)
7.1. pn-junction in an external voltage 589

Substituting this into (7.37), we find

Dn
j n ( z p ) = enp- (eeuBjkT - 1) coth[(b - z p ) / L n ] . (7.40)
Ln

The current ig in the base circuit would be identical with the emitter current
iE, and both currents would equal the current through the emitter-base pn-
junction, if the base were to be infinitely expanded. For finite base width,
however, the emitter current differs from the base current because the latter
only takes contributions from the portion of the electrons injected into the
base which also recombine in the base. We denote the pertinent current
density as jnT. The first term of formula (7.23) expresses it as

(7.41)

The upper boundary of the recombination region, which in (7.23) is located


at infinity, is replaced here by the finite base width b. Furthermore, xn and
x p are interchanged. The recombination rate R(x) in (7.41) can, as before, be
calculated from the minority charge carrier concentration n ( z )= np+An(z)
in the p-region by means of relation (7.24). In this, expression (7.39) has to
be used for An(x), yielding

The base current ig, like the emitter current in (7.36), also involves the
injection current j p ( z n ) Aof holes from the base into the emitter, besides
the electron current j,A. Altogether, we therefore have

The collector current ic need not be calculated separately, because it is al-


ready determined by Kirchhoff's law (7.35). The main source of this current
are electrons which leave the emitter and do not recombine in the base but
diffuse further to the collector where they are sucked up. This current contri-
bution is entirely due to the coupling between the collector-base pn-junction
and the emitter-base np-junction; it would not appear at all at a single,
separate pn-junction. Another current which flows through the collector-
base pn-junction is the true pn-current, which represents the current which
would exist if this junction were not coupled to a second one. However, the
collector-basepn-junction is biased in blocking direction, so that the true pn-
current is an extraction or generation current, which because of its smallness
590 Chapter 7. Semiconductor junctions under non-equilibrium conditions

Figure 7.9: Characteristics of a bipolar transistor made of Si in the common-


emitter configuration. (After Volz, 1986.)

may be completely neglected. This approximation was used above tacitly,


as we identified the current through the emitter-base pn-junction with the
true pn-current of this junction, without adding the hole extraction current
flowing in from the collector-base pn-junction on the right-hand side .
With this, the three currents i E , i g , ic we sought are fully determined.
In Figure 7.9 some characteristic dependencies for i g and ic are shown
in the case of a npn-transistor made of Si. We can now proceed to the
question of what conditions are needed for the bipolar transistor to operate
as an amplifier. As we will see below, significant amplification by a bipolar
transistor results when the base current is only a small portion of the emitter
current. In order for that to be the case, on the one hand, j,,A must be
small. Considering (7.42)) this means that the base width b must be small
in comparison with the diffusion length L , of the minority charge carriers.
On the other hand, the hole injection current j p ( z , ) A is not allowed to be
too large. This can be achieved by low doping of the base in comparison
with the emitter, because jp(z,) is proportional to p, (as may be seen from
(7.38)) and j n ( z p )is proportional to np (see relation 7.39). If the doping
n, of the emitter is substantially higher than the doping p, of the base, i.e.
if nn >> p, holds, then it follows from the mass action law that p, << np.
The hole injection current j p ( z n ) Afrom the base is therefore actually small
compared to the electron emission current j n ( z p ) Afrom the emitter. For
simplicity, we neglect it completely in what follows.
Suppose that the potential p~ of the base is changed by dpg, while
the potentials p~ and p c of the emitter and collector are kept constant.
7.1. pn-junction in an external voltage 591

Correspondingly, the voltage U B in the base circuit changes by d u B = d p B ,


while the voltage U D in the collector circuit remains constant. The currents,
however, will change in both circuits in accordance with

(2)~~ (2%) = cosh[(b - z p ) / L n ] d u g uc ,


(7.45)

The current amplification 1,3 is defined as the ratio of the current change in
the collector circuit to the current change in the base circuit. From (7.44)
and (7.45) it follows that

(7.46)

Here, the extent x p of the space charge region in the base has been neglected
in comparison with the entire base width b. If b << L,, a current amplifi-
cation occurs according to relation (7.46) in the sense that a small current
change in the base circuit leads to a large current change in the collector
circuit.
What is important, however, is not primarily amplification of current,
but amplification of electric power. Initially, one might think that power
amplification would be determined by the ratio of the power change in the
collector circuit to the power change in the base circuit. This ratio would
have, however, no value because in the switching scheme of Figure 7.7 all
the power of the collector circuit is transferred to heat. For electric power
to be useful, a working element must be included in the collector circuit,
say a resistor RL. The voltage drop across it is ~ c R L .Taken together with
the voltage between the emitter and the collector of the transistor, the total
+
voltage in the collector circuit is U c icRL. The transistor resistance is ap-
proximately given by the resistance of the blocked pn-junction between base
and collector. As such, it is practically independent of the base voltage. We
consider, in particular, the case in which R L equals the collector resistance of
the transistor. That means that the voltage drop across the working resistor
R L is the same as the voltage U c between the collector and emitter. What
we have to calculate are the power changes d P L in the working resistor R L
and d P B in the base circuit, if the base voltage U B changes by d U B while
keeping the collector-emitter voltage at a constant value U c . We have

(7.47)
592 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

Figure 7.10: Structure of a bipolar Aluminum


npn-transistor with a buried collec- / I \ SiOD

tor.

.......... ..........
.......... n-Silicon ..........
..........
.......... :::::::::A
I/.._.J ............ ...........
..........................................
n*-Silicon
p3iG-I

(7.48)

The ratio of the first term in the square bracket of (7.48) to the second is
kT to U g . Under normal conditions this is small, and neglecting it the ratio
of the power changes dPL and dPB takes the simple form

(Z) = P UUC
B. (7.49)

Therefore, the power amplification turns out to be proportional to the cur-


rent amplification. As far as the proportionality factor is concerned, we
note that the voltage drop across the collector-base np-junction (biased in
blocking direction), is larger than the voltage U B across the baseemitter
np-junction (biased in flow direction). Because of this, the collector-emitter
voltage U c is larger than U g , and the the power amplification is in fact
larger than the current amplification. Small changes of electric power in the
base circuit are therefore correlated with large power changes in the working
resistor. Of course, this does not mean amplification of power in the sense
that a large power is generated by a small power. Rather, it means that a
large useful power is tuned by a small control power.
The base circuit is also called the input circuit, and the collector circuit
the output circuit. Using this terminology one can also say that the bipolar
transistor allows the control of a large output power by means of a small
input power. The input resistance of the bipolar transistor, i.e. that of the
base circuit, is relatively low, since the corresponding pn-junction is biased
in the flow direction. The output resistance, i.e. that of the collector circuit,
is relatively large compared to the former, because the second pn-junction
operates in the blocking direction. This particular feature of the bipolar
transistor essentially determines its application in electronic circuits. In
the case of an unipolar transistor like the MOSFET, the relationships are
7.1. pn-junction in an external voltage 593

P I n n

Figure 7.11: Lineup of the valence and conduction band edges at a pn-junction
between moderately doped regions (a) and heavily doped regions (b).

reversed, as we will see below; there the input resistance is high and the
output resistance low.
Due to the planar fabrication technology of microelectronics, the bipolar
transistor, as it is actually used in integrated circuits, is structured differently
than the one used in our theoretical analysis. A more realistic structure is
shown in Figure 7.10.

7.1.6 Tunnel diode

The injection-recombination mechanism of current flow in a pn-junction is


valid as long as the doping of the two regions is not extremely heavy ~

strictly speaking, as long as the bulk Fermi levels Epn and EF, do not lie in
the respective bands (Figure 7.11a). When the latter occurs (Figure 7.11b),
then there are energy levels at the conduction band bottom which are at
the same position as some of the energy levels at the valence band top. The
two energy bands partially overlap. In the overlap region each energy level
belongs simultaneously to valence and conduction band states. However,
the corresponding wavefunctions are localized in different regions of the pn-
junction - those of conduction band states in the n-region, and those of
valence band states in the pregion. An electron from a conduction band
state with an energy level in the overlap region can transfer into a valence
band state at the same energy. One then says that the electron tunnels from
the conduction band into the valence band (also see section 3.8 on tunneling
due to an external electric field). Analogously, electrons from the valence
band can tunnel into the conduction band. The probability for a tunneling
transition is the same in both directions.
594 Chapter 7. Semiconductor junctions under non-equilibrium conditions

Figure 7.12: Current-voltage char-


I
acteristics of a tunnel diode.

).
0 I
0 02 084
u/v-

In order for an electron of a given energy to actually be able to tunnel,


it must find unoccupied states of the same energy in the other band. In
equilibrium at temperature T = 0, the only possibility would be electrons
having energies above the Fermi level, but there are none such. This means
that in equilibrium at T = 0 no current can flow due to tunneling transitions
between the valence and conduction bands. This equilibrium result also
remains valid for T > 0. For simplicity, we proceed with the case T = 0
below. Applying a weak positive voltage to the pn-junction, the valence band
in the p-region is lowered relative to the conduction band in the n-region,
and the same happens with the Fermi levels in the two regions. Then there
are electrons of the conduction band in the n-region that can find unoccupied
states of the same energy in the valence band on the p-side of the junction.
Therefore, an electron tunneling current flows from the n-region into the
p-region of the junction. Further increasing the positive voltage, a point
is reached at which the valence band top in the p-region drops below the
conduction band bottom in the n-region. Then there are no more common
energy levels on the two sides of the junction, and tunneling is no longer
possible. The tunneling current through the junction vanishes. pn-junctions
with the above described properties are called tunnel or Esaki diodes.
In Figure 7.12 the current-voltage characteristic of a tunnel diode is
shown schematically. The dashed curve marks the tunneling current con-
tribution to the total current which also contains, besides the tunneling
current, a drift-diffusion current. Moreover, tunneling can resume with the
aid of phonons, so that an energy difference between the tunneling states on
the two sides of the junction is allowed when bridged by the absorption or
emission of phonons. This process is referred to as phonon-assisted tunneling
(in analogy to photon-assisted tunneling treated in section 3.8).
7.2. pn-junction in interaction with light 595

7.2 pn-junction in interaction with light

7.2.1 Photoeffect at a pn-junction. Photodiode and photo-


voltaic element
If a homogeneous semiconductor sample is irradiated with light, electron-hole
pairs are generated. Only part of them is removed by recombination, so that
in the stationary state there remain excess concentrations An of electrons
and A p of holes. These may be calculated from the pair of equations (5.11))
(5.12) which state the equality of the generation rate g and recombination
rate R in the present case. With R of (5.68) and g of (5.27), it follows that

An = Ap = I-g = I-
(E) (7.50)

where I- is the recombination lifetime of the electron-hole pairs and Tzw is the
photon energy of the radiation (the latter is taken to be monochromatic),
Using equations (7.11) and (7.12), the corresponding quasi Fermi levels p n
and pP of, respectively, electrons and holes may be calculated. Due to the
optical excitation they have different values - that of the electrons is raised
and that of the holes is lowered in comparison with the common equilibrium
value. If a voltage is applied to the sample, the current flow is stronger
when the sample is illuminated than without illumination. This phenomenon
is called the internal photoeffect. Alternatively, if one keeps the current
constant, then the photoeffect lowers the voltage drop across the sample.
The illumination leads, so to say, to a negative pre-voltage of the sample.
The latter is just the difference of the quasi Fermi levels of the electrons and
holes.
In these considerations, a spatially homogeneous distribution of light in-
tensity within the semiconductor sample has been assumed. However, since
the light is absorbed in the excitation of the electron-hole pairs, this assump-
tion is not justified. The intensity decays exponentially in the direction of
light propagation. This leads to a corresponding spatial inhomogeneity of
the electron and hole concentrations, which gives rise to diffusion. If the dif-
fusion coefficients of electrons and holes are the same, then diffusion makes
no contribution to electric charge transport since the electron and hole cur-
rents exactly compensate each other. If the diffusion coefficients are different,
however, as often actually happens, either a net current or voltage will arise
in the light propagation direction, depending on whether the two ends of
the sample are electrically connected or not. This phenomenon is called the
Dember effect.
The homogeneous semiconductor photoeffects discussed above are rela-
tively small in comparison with the photoeffects which occur at semicon-
596 Chapter 7. Semiconductor junctions under non-equilibrium conditions

ductor junctions with reverse biased space charge regions. As examples of


the latter, we have the depletion region at a pn-junction, also at a metal-
semiconductor junction, an insulator-semiconductor junction or a free sur-
face. In the case of a pn-junction, various configurations are possible: the
light can propagate parallel to the boundary plane between the p- and n-
regions or normal to it, and in the latter case, the depletion region can be
smaller or larger than the penetration depth of light, etc.
Below, we will deal with a pn-junction and assume that the light prop-
agates perpendicular to the boundary plane between the p- and n-regions,
(parallel to the x-axis of our Cartesian coordinate system, see Figure 7.13).
The corresponding pn-structure is, contrary to what has been assumed thus
far, no longer infinitely extended in both directions but bounded in the pos-
itive z-direction by a plane surface at 2 = 20. We take this surface to lie
on the n-side of the junction, far enough from the depletion region so that
the light is completely absorbed before reaching this region. This assump-
tion simplifies our theoretical description, although it does not essentially
influence its results. Under the assumed conditions, the optically excited
electrons and holes diffuse away from the surface, towards the depletion re-
gion. While the holes are attracted from the negatively charged p-side of
this region, the electrons are repelled from there (see Figure 7.13). In the
space charge field of the depleted p-region the holes drift deeper into the
sample where they recombine. The electrons stay in the surface surface re-
gion and recombine there. Consequently, a net hole current &hob flows from
the surface into the bulk. Its magnitude is determined by the hole diffusion
current at the boundary x = x,, of the depletion region. The maximum pos-
sible value j p h t o of the hole current is given by the photon current density
( I l h w ) multiplied by -el i.e.

(7.51)

This upper limit is reached if all optically excited holes diffuse as far as the
plane 3: = xn, which means that they do not recombine, neither at the surface
nor in the bulk of the sample. We assume this condition to be fulfilled.
The flow of positive charge through the np-junction from its n- to its p-
side, is just the opposite of what happened at the np-junction in establishing
equilibrium without illumination by light. In the latter case, the transfer of
charge results in the diffusion voltage. Therefore, exposure to light reduces
the diffusion voltage. The voltage change U$ooto is called the photovoltage. In
the case of the open pn-circuit considered here, in which the outer planes of
the p- and n-regions are not electrically connected, the total current density
jgbl in a stationary state must vanish. The latter is composed not only of
the photocurrent density, jphoto, but also of the current density j through the
pn-junction under the photovoltage U$ooto. This voltage takes a value which
7.2. pn-junction in interaction with light 597

Figure 7.13: pn-junction subject to light illumination.

Figure 7.14: Electric circuit with an illuminated pn-junction.

guarantees the vanishing of the total current density jgtal.Using expression


(7.28) for the current through the pn-junction, it follows that

(7.52)

If the outer planes of the p- and n-regions are electrically connected (see
Figure 7.14), and if the resistance R in this circuit is 0 (short-circuit), then
the voltage drop across the pn-junction is also zero, and the current

tota
.O at = iphoto. (7.53)
flows. If the resistance in the pn-circuit is neither 0 nor infinity, but has
a finite value R , then the photovoltage takes a smaller value UpRhoto than
Uphoto, and the total current density takes a smaller value jEtal than j:oM.
The two quantities jEbl and Ugot0 are determined by the equations

(7.54)
598 Chapter 7. Semiconductor junctions under non-equilibrium conditions

Here, A is the illuminated area of the pn-junction. Depending on the size


of the working resistor R , the pn-junction can either be used to transform
small light-signals into voltage signals, i.e. as a photodetector, or for the
generation of electric power, i.e. as a photovoltaic element.

P hot odet ect or

In the case of the photodetector, the working resistance R must be chosen


as large as possible, so that the largest possible photovoltage occurs. For
R = 00 it follows from equations (7.54) and (7.55) that

kT
(7.56)

This relation shows that Urhto is more sensitive to the light intensity I for
smaller saturation current density j, of the pn-junction. In subsection 7.1.4,
j, was estimated to be 10-"A/cm2 for a pn-junction made of Si. From this,
one can conclude that for hw = 1 e V , a light power of 10-l' W / c m 2 will
create a photovoltage of the order of magnitude 10 mV. The upper limit of
photovoltage is given by the diffusion voltage U D . This cannot be surpassed
because the space charge region would cease to exist if Upmbto were to exceed
Uo. In the latter case, no separation of optically excited electron-hole pairs
could take place, which is crucial for operation of the device.
Photodetectors involving semiconductor junctions are referred to as photodi-
odes. Beside the pn-photodiode discussed above one has also pin-photodiodes
with particularly wide depletion regions (symbolized by the 'i' in pin). Other
examples are met al-semiconductor or heteroj unction photodiodes.

Photovoltaic element

In the case of the photodetector, the usable electric power is very small and,
in fact, for R = 00 it vanishes. If the illuminated pn-junction is to produce
electric power, then R cannot be made infinitely large. On the other hand,
R also cannot be taken arbitrarily small, because then there would be no
voltage drop at the pn-junction at all and the power would vanish for this
reason. For finite values of R , a non-vanishing electric power is generated,
with light energy converted directly into electric energy. The useful electric
power R(Ajztd)2 can be calculated by solving equations (7.54) and (7.55)
with respect to jzhland Upbto. The calculation shows that the efficiency of
photovoltaic energy conversion can, theoretically, reach up to about 40 %.
In practice, it is not much more than 10 % in most cases currently. Photo-
voltaic elements have wide and important practical application as solar cells,
particularly those made of silicon.
7.2. pn-junction in interaction with light 599

7.2.2 Laser diode


If a pn-junction is biased in flow direction, the minority carriers diffuse into
the depth of the two bulk regions, where they recombine. The manner in
which recombination proceeds is unimportant if one is only concerned with
current transport. In our previous calculations we assumed radiationless
recombination through deep centers (Shockley-Read-Hall recombination).
This was justified inasmuch as the recombination rate RmT of this process
appreciably exceeds that of other recombination processes in many cases.
Under appropriate conditions, however, radiative recombination processes
can also play an important role. In such processes electron-hole pairs anni-
hilate with the emission of photons (see section 5.2). If the emission proceeds
spontaneously, then the recombination rate Rap, is given by the expression

(7.57)

where rSwT is the corresponding spontaneous radiative recombination life


time and An = n - no is the excess concentration of electrons beyond the
equilibrium value no. On the other hand, light emission can be induced by
the radiation itself. Then one speaks of stimulated emission processes. In
the following considerations we assume monochromatic light with a definite
phase and propagation direction. Its photon energy hw is taken to match
the band gap energy of the semiconductor. The light propagation is de-
scribed by the photon current density S (flux of photons crossing unit area
in unit time), which follows from the light intensity I by means of the rela-
tion I = hwS. In terms of S, the radiative recombination rate RstimTdue to
stimulated emission may be written as

(7.58)
where g is the so-called gain coefficient. The latter is defined as the negative
of the absorption coefficient, strictly speaking, that part of the absorption
coefficient which is due to the excitation of electrons from the valence into
the conduction band. The dimension of g is, therefore, that of a reciprocal
length.
In order for g to be greater than zero, the semiconductor must be put into
an excited state. The energy of this excited state (relative to that of the
ground state) is the energy emitted in stimulated emission when irradiating
the semiconductor with light. If the emitted radiation is to prevail over that
absorbed, the number of stimulated transitions from the excited state to the
ground state must be larger than the number of transitions in the opposite
direction. This case is realized when, in the excited state, the conduction
band edge is occupied by more electrons than the valence band edge. This
is called an occupation inversion In the case of equal occupation of the two
600 Chapter 7. Semiconductor junctions under non-equilibrium conditions

band edges, absorption and stimulated emissions compensate each other,


which means that the semiconductor is transparent (not counting sponta-
neously emitted photons and the losses by absorption processes which are
not related to transitions of electrons from the valence to the conduction
band, like absorption by free carriers or phonons). The gain coefficient g
is zero in the transparent state. Generally, in accordance with the remarks
above, g is a function g(n) of the electron concentration n in the conduc-
tion band. The particular concentration which corresponds to the case of
transparency g ( n ) = 0, is called the transparency concentration nt,. In the
vicinity of nt,, g ( n ) may be represented by a linear Taylor expansion with
respect to ( n - ntr),such that

9= d ( n - ntr), (7.59)

where g is a constant. Although this expansion holds for both signs of


( n - nt,), we assume ( n - nt,) > 0 below. .This is to say that we consider
only cases in which stimulated emission dominates over absorption due to
optical transitions from the valence to the conduction band.
In order to achieve occupation inversion, heavily doped pn-junctions are
required, for which the bulk Fermi levels of the p- and n-regions lie, respec-
tively, in the valence or conduction bands (see Figure 7.15). By applying
a sufficiently large flow voltage to such a junction, a non-equilibrium state
can be created in which, close to the nominal transition region at x = 0,
the two quasi Fermi levels differ by more than the energy gap. This indi-
cates occupation inversion according to the definition above. The layer of
the pn-junction in which inversion occurs is termed the active region There,
the first of the two laser conditions, the occurrence of occupation inversion,
is fulfilled. To also meet the second laser condition, namely the presence
of radiation feedback, one has to provide that the active region operates as
a Fabry-Perot resonator in one of the two directions perpendicular to the
x-axis (here we take the y-direction, see Figure 7.16). This can easily be
achieved by simply cleaving the semiconductor sample perpendicular to the
y-axis. The cleavage planes function as mirror planes because of the high
refractive index of semiconductor materials. A pn-junction prepared in this
way is called a laser diode.
The useful laser output is the radiation which passes out through the
resonator planes, as they are not ideal mirrors. This radiation is completely
determined by the radiation within the resonator. Thus, the latter must be
calculated to understand the laser diode quantitatively, and we will do this
below on the basis of the qualitative considerations above.
As we have seen, occupation inversion at thepn-junction comes about by
the injection of minority charge carriers under the influence of a flow voltage.
In section 7.1, the corresponding injection current density j was determined
7.2. pn-junction in interaction with light 601

Figure 7.15: Heavily doped pn-junction without voltage (a) and with voltage in
flow direction (b). In case (b) occupation inversion occurs in the vicinity of the
pn-boundary at x = 0.

as a function of applied voltage. Here, we need the relation between j and the
photon current density S, taking the electric current density j be given. As
we shall see, the functional dependence of S ( j ) can be calculated just from
the continuity equations for the carrier currents j,, j , and the photon current
density S. For simplicity, we assume a symmetric pn-junction, for which it

Figure 7.16: Geometry of a laser diode.


602 Chapter 7. Semiconductor junctions under non-equilibrium conditions

suffices to consider only the electron current density j,. Also, the active
region is symmetric with respect to x = 0 under this assumption, so that
it extends between -(1/2)x,4 and ( 1 / 2 ) z ~with ;CA being the total length
of the region. The continuity equation for the electron current density j ,
was written down in Chapter 5 in general form. The derived relation (5.20)
will be specialized here in the following way. First of all, we recall that the
total current density j = j n ( z ) +jp(z)is spatially constant in the stationary
state, while the two components j n ( x ) and &,(I) vary with I. On the right-
hand side of the active region, j n ( z ) is almost identical with j , while &(z)
practically vanishes. On the left-hand side of the active region, jp(x) has
almost the same value as j , while jn(x) is close to zero. Consequently, in
stationary state, we find for the active region the approximate relation

(7.60)

Relaxation processes which change the electron concentration in time include


non-radiative recombination as well as spontaneous radiative and induced
radiative recombinations. For the assumed symmetric pn-junction, the non-
radiative recombination rate Rmr is determined by a common lifetime T
for electrons and holes, as follows,

An
R-p= -. (7.61)
7-

Applying expressions (7.57), (7.58) and (5.60), (5.61) to the continuity equa-
tion (5.20) for the electron current, we have

an
---=
j An An
gs. (7.62)
at exA 78pDnT

To formulate the continuity equation for the photon current density S , an


expression for the pertinent photon density is needed. According to electro-
dynamics, this density is (S/c) where c denotes the group velocity of light
propagation. The source density of the photon current is zero since S flows
in y-direction and the pn-junction can be considered as homogeneous in this
direction. The photon density does involve,external sources, however, due to
spontaneous and stimulated radiative annihilation emission processes. Since
each radiatively recombining electron-hole pair creates a photon, the corre-
sponding source density is - apart from a modification which will be detailed
immediately below - given by the sum of the radiative recombination rates,
i.e. it is the negative of the last two terms on the right-hand side of equa-
tion (7.62). The modification affects the second term (An/rSmT),which
describes spontaneous emission. Only a small fraction of the spontaneously
emitted photons have the same propagation direction and phase as those
due to stimulated emission. Thus this term must be multiplied by a factor
7.2. pn-junction in interaction with light 603

,B substantially less than unity if incorporated in the continuity equation for


S. The source terms due to spontaneous and stimulated emission must still
be complemented by an additional term, which accounts for light absorption
processes which are not due to transitions of electrons from the valence to
the conduction band (see above). We denote the corresponding absorption
coefficient by a. The pertinent (negative) photon source density is -as. In
a , one can also include losses due to the penetration of photons through the
resonator planes which are unavoidable because otherwise the laser would be
ineffective for the world outside. Finally, the continuity equation for photon
current density takes the form

(7.63)

where the left-hand side is the change of the photon density per second. We
are interested in the stationary state of the laser defined by

an dS
- -- - = 0. (7.64)
at at
We first consider g to be different from a. As we will see, in stationary state,
this necessarily means that g is less than a:
g<cw
The corresponding photon current density S< follows from equation (7.63)
as

(7.65)

Since S<, An and rSpT are positive quantities, only the case g < a is
meaningful, as indicated above. The photon current density S< given by
equation (7.65) is relatively small because the spontaneous radiative lifetime
T~~~~ has relatively large values and P is small compared to one. This is

true on the proviso that g does not approach the value of a very closely, or,
in other words, as long as the electron concentration n is distinctly smaller
than the threshold value nth defined by the relation

0: = 9/(nttZ - nt7-1. (7.66)


The dependence of the photon current density S< on the carrier current
density j follows from equations (7.62) and (7.63) as

j
s<= (7.67)
ez~[g< + ( l / P ) ( l + rspa~/r)(a- g<)I
Here, g has been replaced by g< to indicate that the value of g has to be taken
as that which follows from g / ( n - ,ntT) for n < nth. If spontaneous emission
604 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

is assumed to be extremely weak, i.e. if rSpar= 00, then it follows that


S< = 0. This result indicates that under the assumption n < nth the photon
current density S< is determined by spontaneous and not by stimulated
emission. The pn-junction does not operate as a laser but as a spontaneous
radiation emitter. Devices which rely on spontaneous emission, the so-called
light emitting diodes (LEDs), are, however, constructed differently than laser
diodes, in particular, they do not have a Fabry-Perot resonator. Also, their
physical basis differs from that of the laser diode. Instead of focusing on the
radiation propagating in one direction with the same phase and frequency,
the total radiation is important. We will not discuss LEDs in greater detail
here.
The photon current density S remains relatively small if the lifetime
rsponrfor spontaneous emission is identified with the value that actually
occurs in semiconductors, instead of replacing it by 00. The contribution of
radiative recombination to the carrier current density j in equation (7.62)
can be neglected under these circumstances, and we have, approximately,

(7.68)

This equation relates the electric current density j and the electron con-
centration n below the laser threshold. The threshold concentration nth
corresponds to a threshold carrier current density j t h as

(7.69)

Laser operation can occur only if the threshold current density j t h is ex-
ceeded. We will assume that this is the case. Then, by definition, we have g
equal to a:

The photon current density can now take non-vanishing values if rVar = 00.
This does not contradict the photon current continuity equation (7.63) under
the stationary conditions (7.64) because (g - a ) S now vanishes since (g - a)
does so. The actual value S> of S for a given current density j > j t h may be
obtained from the carrier continuity equation (7.62). In the approximation
rspar = 00, it yields

(7.70)

We will show that S> is in fact considerably larger than Sc. To this end,
we consider S> given by equation (7.70) for j = 2jth, and examine the ratio
of s>(2jth) to s<of (7.67) for j = j l h , obtaining
7.2. pn-junction in interaction with light 605

3
laser diode as a function of current den-
0
\
sity j (schematically). v)

I / *
Jth j 1a.u.

Generally, the spontaneous radiative lifetime 7spaT is much larger than the
non-radiative lifetime T. One has (7spaT//37) > 1 since ,B << 1. The ratio
( g < / m ) is a positive number measurably smaller than 1. Therefore, the
right-hand side of equation (7.71) is large compared to 1, meaning that
the value of the photon current density S, above the threshold current
density substantially exceeds the value S< below threshold. This is the
manifestation of the laser effect. One can also say that the pn-junction
functions as a laser diode when j > j t h holds. The functional dependence
of the radiation intensity over the entire current density range is shown
in Figure 7.17 schematically. The characteristic feature of the curve is its
steep rise at the threshold density jth. This point marks the transition from
spontaneous to stimulated emission, thus the onset of the laser regime.
To estimate the order of magnitude of the threshold current and of the
radiation intensity of a laser diode, we choose parameters close to those of
G ~ A SWe . take ntr = ~ ,= 2 . 1 0 ~ ' cmv3, X A = 1 p m , r =
c m ~ nth
lo-' s, and a = lo3 cm-'. With these values the threshold current density
of (7.69) is j t h = 103A ern-'. The corresponding current strength amounts
to 1 m A if a contact area of 10 x 10 ,urn2 is assumed. For current strength
2 m A the radiation intensity RwS within the resonator is about 10 mW. This
example shows that the radiation intensities of semiconductor laser diodes
are large enough to be employed for practical purposes, in particular in
information processing and communication. Currently, semiconductor laser
diodes are widely used in optical fiber communication, CD-players, scientific
instruments and in many other ways.
The spontaneously emitting diodes (LEDs) have also attained wide prac-
tical application, e.g., for signal generation and pattern display. Since the
radiation emitted by laser diodes and LEDs lies in the photon energy range
close to the band gap, one can theoretically cover the entire frequency spec-
trum from the infrared to the near UV by a proper choice of semiconductor
606 Chapter 7. Semiconductor junctions under non-equilibrium conditions

material. The practical difficulty of accomplishing this goal increases with


frequency. There are various reasons for this, the most general being the
separation of the operating non-equilibrium state from equilibrium: it be-
comes larger, and therefore more difficult to achieve, if the gap increases.
At this time, usable laser diodes and LEDs exist from the infrared up to the
green spectral region. Blue and ultraviolet light emitting devices are still
the subject of research.

7.3 Metal-semiconductor junction in an external


voltage. Rectifiers
Only non-ohmic metal-semiconductor junctions, i.e. Schottky junctions are
suitable for active electronic functions. Therefore, we restrict our consider-
ations here to the latter. Moreover, we suppose that the semiconductor is
n-type. For p-type semiconductors analogous results may be derived. The
two most important electronic devices based on Schottky contacts are the
metal-semiconductor rectifier and the MESFET. The physical principle of
MESFET operation is similar to that of the MOSFET. It relies on the change
of charge carrier concentration due to an external voltage in a thin layer of
the semiconductor at the interface with, respectively, the metal or insulator.
We will discuss this in section 7.4 in connection with the MOSFET. Here, we
concentrate on the physical principle of operation of the metal-semiconductor
rectifier, involving current transport through a biased Schottky junction.
As in the case of the pn-junction, the application of an external voltage
U to a Schottky junction results in a reduction of the diffusion voltage UD
of (6.58) by U , so that

(7.72)

We will assume that the Schottky approximation can be used in the semi-
conductor, i.e., that the inequality (7.4) is fulfilled. Then the division of the
semiconductor into depletion and bulk regions also retains meaning here.
The width X B of the depletion region follows from (6.62) if U D there is re-
placed by U D - U . The current in the bulk region of the semiconductor
is carried by majority carriers, just as in the case of the pn-junction. For
the assumed metal-n-type semiconductor junction, the majority carriers are
electrons on both sides of the junction. At the interface between the two
materials, electrons are thermionically emitted from the metal into the semi-
conductor, and vice versa, electrons from the semiconductor are emitted into
the metal. We denote the particle current density from the metal by i M S ,
and that from the semiconductor by i S M . The two partial current densities
add up to the total current density i through the junction. All three current
7.3. Metal-semiconductor junction in an external voltage. Rectifiers 607

a) 0 - r 8 u-0
Metol n- Semiconductor

bl
c) +t+ UCO

___.
u70

c
---T'
---I---: L- \
i": I
x x x
I

iI ------
I
I
x x A_-
I
I
I
I
L
XB X

Figure 7.18: Illustration of current flow at a Schottky-contact, a) without voltage,


b) with voltage in flow direction, and c) with voltage in blocking direction. The
crosses in b) and c) indicate that the Fermi level is not defined in the space charge
region.

( U )i ( U ) of the external voltage U ,


densities are functions i ~ s ( U ) , i s ~ and
and
i ( U ) = iMS(U)- iSM(Cr). (7.73)
In the absence of any voltage (see Figure 7.18a) one has i(0)= 0, i.e., the
two partial currents compensate each other, so that
i M S ( 0 ) = 1'SM(O). (7.74)
The application of a voltage U causes the space charge barrier of the semi-
conductor to change, for U > 0 it is lowered (Figure 7.18b), and for U < 0 it
608 Chapter 7. Semiconductorjunctions under non-equilibrium conditions

is increased (Figure 7.18~).Accordingly, the current i s ~ ( Ufrom


) the semi-
conductor into the metal increases (U > 0) or decreases (U < 0). However,
the current from the metal into the semiconductor remains constant since
the potential barrier to be overcome by the electrons, the Schottky barrier
Cpg, does not change with applied voltage. Thus,

iMS(U)= iMS(0). (7.75)

The total current density i in the presence of the voltage U may then be
written as
i ( U ) = i ~ s ( 0-) i s ~ ( U ) . (7.76)
We first calculate i ~ s ( 0 )In. this, we consider a metal electron which, after
passing into the semiconductor, has quasi-wavevector k and energy Ec(k).
Since no momentum and energy change takes place in crossing the interface,
the quasi-wavevector of this electron is k and its energy Ec(k) in the metal as
well. The probability to find an electron in the metal with energy E = Ec(k)
is given by the Fermi distribution f i ~ ( Eof) the metal,

(7.77)

The current density i ~ s ( 0is) the statistical average of the speed component
h-'(Ll/LlkX)Ec(k) in the positive 2-direction, per unit volume. According to
the general rule (5.3) for calculating statistical averages we have

The boundaries of the first B Z were put to infinity here, involving only a
slight error. In fM(Ec(k)) we replace -EFM by the work function @ of the
metal. This also indicates that, henceforth, the energy origin will no longer
be taken at the top of the semiconductor valence band, but at the common
vacuum level of the two materials. Since [Ec(k)+@] >> k T , the Fermi distri-
bution (7.77) can be approximated by the Boltzmann distribution, whence

(7.79)
For Ec(k) we use an isotropic parabolic dispersion law with effective mass
m:. Then relation (7.78) takes the form
7.3. Metal-semiconductor junction in an external voltage. Rectifiers 609

The conduction band edge, referred to the vacuum level, lies at E, = - X


so that @ +E, = CI, - X is equal to the Schottky barrier @B. After a short
calculation we find

i ~ ~ (=0 ) -OB/kT, (7.81)


4

with
(7.82)

as the average speed component in positive z-direction and n as the electron


concentration in the semiconductor. Somewhat surprisingly, through mE and
n quantities characteristic of the semiconductor occur in expression (7.81)
for i ~ s ( Oalthough
), the current originates in the metal and one might have
expected the appearance of characteristic metal quantities only. The reason
is that for U = 0 (which is the case we are considering now) thermodynamic
equilibrium exists, and the current from the metal into the semiconductor
must equal the current from the semiconductor into the metal. That the
latter depends on semiconductor quantities is obvious.
The interpretation of i ~ ~ (as0 a)current i ~ ~ (from0 ) the semiconductor
into the metal in thermodynamic equilibrium may also be used to calculate
i s ~ ( Uin) the presence of a voltage. For U = 0, first of all, this interpretation
yields the expression

If a voltage U is applied to the junction, the conduction band edge of the


semiconductor at the interface with the metal shifts from its equilibrium
position E, to the new position E, - e U . This change in (7.83) leads to the
relation

iSM(U) ,eU/kT Z S M ( 0 ) .
I (7.84)
Correspondingly, the total particle current density i ( U ) is

(7.85)
For the electric charge current density j = -ei it follows that

(7.86)

with
I -@p,/kT
j , = Tenvoe (7.87)
Lf
6 10 Chapter 7. Semiconductor junctions under non-equihbrium conditions

For positive voltages U , the current density j grows exponentially with in-
creasing U , while for negative U of increasing magnitudes, j approaches
the saturation current density -js. The latter occurs when 1UI >> kT holds.
The Schottky junction therefore operates as a rectifier, with positive voltages
corresponding to the flow direction, and negative to the blocking direction.
The following estimate shows that j , can assume quite large values. For
T = 300 K , N D = 10l6 cm-3 and @B = 0.25 e V , we have j , 0.2 A / c m 2 .
From this we may conclude that Schottky junctions are suitable for rec-
tification of relatively strong currents. This distinguishes them from pn-
junctions where the saturation current density is commonly much lower (see
section 7.1). The reason for this difference is that the saturation current of
a pn-junction is due to minority carriers, while the saturation current of a
Schottky junction is caused by majority carriers.
The conditions of validity for the current-voltage characteristic of a metal-
semiconductor junction derived above will now be examined. The applica-
tion of formula (7.83) to the calculation of the current density iSM(0) reflects
the assumption that thermionic emission of electrons from the semiconduc-
tor into the metal proceeds unhindered. To appreciate the significance of
this assumption one must first recognize that the emitted electrons originate
in the bulk region of the semiconductor, for in the depletion region none are
available. Unhindered emission can only occur, therefore, if the electrons,
during their flight through the depletion region suffer no collisions. That is
assured if the mean free path length Zf is larger than the space charge width
"B,
lf > X B . (7.88)
For practical Schottky junctions, operating in blocking direction, this con-
dition is often fulfilled, provided the blocking voltage is not too large. For
flow voltages of sufficient magnitude it is always correct. Condition (7.88)
excludes the possibility that the depletion region can even approximately
be in a local equilibrium state. In particular, chemical and electrochemi-
cal potentials cannot be meaningfully defined, not even in a local or 'quasi'
sense. This implies that an essential requirement for the applicability of
the phenomenological equations (5.14) and (5.15) for the current densities is
no longer valid. If the inequality (7.88) is satisfied, drift and diffusion lose
their meaning as transport mechanisms in the depletion region of a metal-
semiconductor junction. The transport proceeds by electrons flying through
the depletion zone unimpeded, which is termed ballistic transport. If, instead
of l f > ZB, the condition
lf << X g (7.89)
holds, then equation (7.83) is no longer valid for use in calculating i S M ( 0 )
and i s ~ ( U ) The
. magnitude relation (7.89) is just the upper part of the
general condition (5.9) for the validity of the phenomenological equations,
7.3. Metal-semiconductor junction in an external voltage. Rectifiers 611

Figure 7.19: Population and depopulation of interface states at an insulator n-type


semiconductor junction.

applied to the Schottky junction. If the condition (7.89) is satisfied, these


equations can be used to calculate i ~ ~ ( 0 In ) . doing so one obtains the
so-called diffusion theory of current flow through the Schottky junction. A
detailed outline of this theory will not be given here. Qualitatively it yields
results similar to the above theory of i s ~ ( U )which
, in this context is called
the thermionic emission theory. In fact, the two cases (7.88) and (7.89) must
also be distinguished in the theory of current flow through a pn-junction.
Starting from the phenomenological current equations in section 7.2, we also
assumed the validity of a condition similar to (7.89) when we considered the
case of a pn-junction. This means that the diffusion theory of current flow
was developed, and the diode theory was omitted, in our earlier discussion
of the pn-junction. For the pn-junction this is justified inasmuch as the
diffusion theory is applicable in most cases, while the thermionic emission
theory is not.
If the depletion region of the metal-semiconductor junction becomes so nar-
612 Chapter 7. Semiconductor junctions under non-equilibrium conditions

row that the electrons can tunnel through the potential barrier, then even
the thermionic emission theory is no longer valid. This occurs in heavily
doped semiconductors. In this case the Schottky contact conducts well in
both directions, and it becomes an ohmic contact. This effect is exploited
in manufacturing contacts in silicon devices. The frequently used contact-
metal, aluminum, would in fact result in a non-ohmic Si-metal contact if the
doping were not sufficiently high. The heavy doping of Si makes this contact
ohmic, as is necessary for the proper operation of devices.

7.4 Insulator-semiconductor junction in an exter-


nal voltage
In Chapter 6 we saw that the properties of an insulator-semiconductor junc-
tion in thermodynamic equilibrium depend on whether interface states are
present or not. If such states do exist, the bands at the interface are already
bent without an applied voltage, and one has a depletion or accumulation
layer there. These arise in consequence of the spatial redistribution of free
charge carriers in the presence of interface states - some of the carriers are
either captured or generated by these states (see Figure 7.19). The semi-
conductor as a whole thereby remains electrically neutral, of course. This
changes if a voltage is applied.

7.4.1 Field effect


Consider an insulator layer of finite thickness d on top of a semi-infinite
semiconductor of either n- or p-type (see Figure 7.20). To applying voltage,
a metal layer must be deposited on the surface of the insulator at x = -d. I f
this layer is put at potential p ( - d ) , and the outer semiconductor boundary
at x = 00 is at potential p(00), then the voltage U across the structure is
given by the relation
U = p ( - d ) - p(00). (7.90)
The electric field strength deep in the semiconductor, i.e. at x = 00, contin-
ues to be zero even in the presence of the external voltage, whence

(7.91)

From electrostatics it follows that under these circumstances, electric charge


is induced in the semiconductor. The amount of this charge turns out to be
the negative of the charge which must be put on the metal layer to establish
the voltage U across the insulator-semiconductor junction. This applies,
strictly speaking, to the case in which the static dielectric constants E I and
E of the insulator and semiconductor, respectively, are equal. If E I differs
7.4. Insulator-semiconductorjunction in an external voltage 613

u<o

UO

Figure 7.20: Formation of depletion and accumulation layers at an insulator-


semiconductor junction subject to an applied voltage.

from t, then the charge on the metal contact has to be multiplied by the
factor ( I / ) to get the induced charge. Since, in general, E I is smaller than
t (ferroelectrics are an exception), the induced charge is generally a fraction
of the charge placed on the metal contact. Inducing charge at an insulator-
semiconductor junction by applying a voltage is also referred to as field effect.
This effect has played an important role in the development of semiconductor
physics. The thought of exploiting it to induce a change of semiconductor
resistance by means of an external voltage, hence, to develop a kind of
a transistor, is obvious and was followed by several scientists in the late
twenties and thirties of this century. Initially, the efforts were not successful,
however. Instead of the field effect transistor, the bipolar transistor was
invented in 1949 - in a sense as a byproduct of the search for the former.
The main reason for the failure of the early work on the field effect transistor
614 Chapter 7. Semiconductor junctions under non-equilibrium conditions

is the existence of interface states at the junction of the insulator and semi-
conductor. If such states are present in sufficiently high density, which was
in fact the case in the early work, all of the induced carriers are captured
by them. Since carrier mobility in these states is zero, the resistance of the
semiconductor does not change. Generally, the number of electrons cap-
tured depends on how many unoccupied interface states are pushed below
the Fermi level by applying the external voltage (see Figure 7.19c), or how
many previously occupied states are lifted above the Fermi level (see Figure
7.19b). The amount of captured charge represents, in this way, a measure of
the number of interface states in an energy interval at the Fermi level of the
size of the applied voltage. By measuring this charge as a function of applied
voltage U , which can be done, e.g., by measuring the capacitance change of
the insulator-semiconductor junction with U , one may obtain experimental
data about the density of interface states at different energies.

7.4.2 Inversion layers


For a field effect transistor to work well, the density of interface states
must be as small as possible. In the following, we will neglect these states
completely and will study the effect of an applied voltage on an insulator-
semiconductor junction under these circumstances. Qualitatively, it can be
ascertained immediately, that for not too large voltages U , an accumulation
or depletion layer is formed depending on the sign of U and the type of
semiconductor (see Figure 7.20). Below, we assume a p-type semiconductor.
Starting from U = 0, an increase of U bends the bands of the semiconductor
at the interface with the insulator downward. Increasing U further, sooner
or later, a point is reached at which the conduction band energy Eg - ecp(0)
at the interface dips down below the Fermi level EF (see Figure 7.21). This
happens if the voltage drop across the semiconductor Us = cp(0) - cp(00) is
larger than (Eg- E F ) / e 3 Cro, i.e. if

(7.92)

holds. Increasing the voltage U further, the conduction band edge E,-ecp(O)
at x = 0 is pushed further down, and the region of the semiconductor in
which the conduction band edge lies below the Fermi level extends to the
right to a point xi > 0 given by the relation

E g - ecp(zi) = E F . (7.93)

(see Figure 7.21). In this region the free charge carriers are not predomi-
nantly holes, as otherwise in a p-type semiconductor, but rather, they are
electrons. The conduction carrier type is inverted in this region, which is
therefore called an inversion layer. The electrons in the inversion layer of a
7.4. Insulator-semiconductorjunction in an external voltage 615

Figure 7.21: Inversion layer at an insulator-ptypesemiconductor junction.

psemiconductor form a degenerate electron gas. Their concentration is of


the order of' magnitude of the effective density of states Nc of the conduction
band. This is much larger than the concentration of holes in the p-type semi-
conductor which hosts the inversion layer, provided the doping of the latter
is not too strong which means as long N A << Nc holds. This requirement
will be assumed to be satisfied here. In order that inversion occur at all, the
voltage drop Us across the semiconductor must obey condition (7.92), which
is also assumed to be valid.
We will now examine the spatial variations of the charge carrier con-
centrations ).( and p ( z ) in the semiconductor, which are determined by
the Fermi distribution f ( E ) of equation (4.13) with the replacement of E F
by the spatially varying chemical potential p ( z ) . The latter follows from
p ( z ) - ecp(z) = E F as
p(.) +
= E F ec p ( X ) . (7.94)
With this expression, the appropriate position-dependent Fermi distribution
f ( E ,x) is

(7.95)

If the Fermi energy E F lies below the conduction band edge, then the ma-
jority of electrons in this band have energy values at and a few kT above the
band edge E, - ecp(x). As long as [E, - ecp(z) - E F ]is substantially larger
616 Chapter 7. Semiconductor junctions under non-equilibrium conditions

than kT,f ( E ,x ) is approximately zero for the energy values mentioned. If,
on the other hand, the Fermi level lies above the conduction band edge,
i.e., if the expression [E, - ecp(x) - EF] is negative and its absolute value is
simultaneously much larger than k T , then f(E, x ) is approximately 1. The
energy interval in which f(E,x ) is neither 1 nor 0 decreases, as compared to
the whole potential interval eU, spanned by the potential e p ( x ) in the semi-
conductor, as eU, increases in comparison with k T . We therefore suppose
that
k T << eU, (7.96)
holds. In a sense, this relation represents a generalization of condition (6.35)
for the validity of the Schottky approximation. Then, considering (7.95) for
energy values of the conduction band, we can approximately write

f(E,X ) = B [ E F - E + ecp(x)] E > E,. (7.97)

Substituting this expression in equation (4.67) for the electron concentration


n ( x ) , we have

Furthermore, beyond the inversion boundary x i , the valence band edge


-ecp(x) initially still lies far below the Fermi energy. Thus, for energy val-
ues E < 0 within the valence band, f(E)= 1, provided x does not lie
too far to the right of x i . The hole concentration p ( x ) therefore vanishes
for not-too-large values of x . At very large x , however, the valence band
edge -ecp(x) approaches the Fermi energy so closely that, approximately,
the same conditions exist as in the bulk. There, p = NA holds if completely
ionized acceptors are assumed, as we do. In the Schottky approximation, the
transition between the two z-regions occurs abruptly, at a point of transition
x p of the depletion region which has yet to be determined. Thus, the hole
concentration p ( x ) , is approximately given by

(7.99)

Evaluating the energy integral for n ( x ) in (7.98) with the square-root ex-
pression (4.44) for the density of states p:&&(E) x p;"'(E), a concentration
+
n ( z )follows which is proportional to [ e p ( x ) EF - E g ] 3 / 2 .Substituting this
into the Poisson equation (5.24), the solution leads to inverse elliptic func-
tions which are difficult to handle analytically. We therefore replace the true
density of states by a model state density Fc which is constant in the energy
7.4. Insulator-semiconductor junction in an external voltage 617

+
interval between E, and EF ecp(x), with the constant 7, of the model de-
termined as follows: We form the average value of the true square-root-like
density of states p?(E) in the occupied energy interval at x = 0, i.e. be-
+
tween Eg and EF eUs. If we were to use this average value for the density
of states throughout the entire inversion layer without any correction, then
the resulting average electron concentration in the inversion layer would be
markedly larger than it really is. Therefore, a correction factor y having a
value between 0 and 1must be introduced in consideration of the decrease of
the upper limit of the averaging interval with increasing I . Thus, we make
the substitution

(7.100)

Using this constant density of states, the electron concentration n(x) in the
inversion layer may be rewritten in the form

where
(7.102)

is the effective density of states of our model. The substitution (7.100) has
the consequence that the true density of states, and with that also the true
electron concentration, is underestimated in the left part of the inversion
layer, and overestimated in the right part, provided that a reasonable choice
of y has been made. Because of this, the potential profile p(x) of the in-
version layer, whose calculation we will now discuss, has a curvature in our
model which differs somewhat from that in reality. Qualitatively, however,
no significant differences occur. For a quasi 2-dimensional electron system,
the results of our model even apply exactly, since in this case the density of
states is constant from the outset.
With the resulting concentrations of free charge carriers as derived above,
the Poisson equation (5.24) in the semiconductor takes the form

IO1 -d _< I I:0,


(7.103)

The solution of this equation must satisfy the boundary conditions (7.90) and
) 0
(7.91). The arbitrary constant of the potential is chosen such that ( ~ ( 0 0=
618 Chapter 7. Semiconductor junctions under non-equilibrium conditions

holds. From the Poisson equation it immediately follows that the potential
and the field strength also vanish at x = xp, i.e. that

V b P ) = 0, (7.104)

(7.105)
The field strength in the insulator is spatially constant and we denote it by
FJ. Then, in particular, we have

(7.106)

For the potential p(x) at x = -d, relation (7.90) and p(00) = 0 yield the
value
p(-d) = U . (7.107)
The four conditions (7.104) to (7.107), as well as the continuity condition
for p(x) and (dpldx) at x = 0 and x = xi, are sufficient to determine
the potential profile p(x) uniquely, along with the three as-yet unspecified
parameters xi,xp and FI. The solution of the Poisson equation under the
above conditions is determined as

u - (X+ d)FI, -d 5 x < 0,


Uo + (Us - Uo)cosh(x/L) - FsL sin(x/L), 0 5 x < xi,
xi i x < Xp,
(7.108)
with
(7.109)

as the characteristic screening length of the electron gas of the inversion

(7.110)
as the electric field strength in the semiconductor at the interface with the
insulator, and
Ua = U - FId (7.111)
as the potential drop across the semiconductor. The inversion boundary xi
and the field strength F obey the equation

tanh(xi/L) = ( U , - Uo)/F,L, (7.112)


which follows from (7.108) for x = zi immediately. Substituting equation
(7.112) into the matching condition for the field strength at xi, we obtain
7.4. Insulator-semiconductorjunction in an external voltage 619

1
cosh(xi/L) [(Us - Uo)2 - (LF,) 2 = =-(xi
NA
N,
- xP)Fs(U, - Uo). (7.113)

By means of (7.112) and (7.113), xi and Fs can be determined as functions


of Us. The value of xi is finite for all Us. We consider here, however, only
the limiting case N A << Fc, in which the right-hand side of (7.113) can be
neglected, so that
Fs = (Us - U o ) / L . (7.114)
Substituting this expression in (7.112), we arrive at the conclusion that the
width xi of the inversion layer should be infinitely large. This is, however,
only an artifact due to the neglect of the depletion region to the right of xi.
In reality, under the condition N A << N,we have only that

xi >> L. (7.115)

The number ns of electrons stored in the inversion layer per cm2 may be
obtained from (7.101) and (7.108) as
-
ns = N,L. (7.116)

With and L taken from, respectively, (7.102) and (7.109), this expression
may be rewritten as

The sheet density ng of electrons in the inversion layer is zero if the voltage
drop Us across the semiconductor equals the threshold voltage Uo marking
the onset of inversion. Above this value, ng has a weak overlinear depen-
dence on Us. Thereby, U , depends on the applied voltage U through the
relation

(7.118)

The ratio of insulator thickness d to the screening length L of the electron


gas of the inversion layer is decisive for the size of the variation of ns as a
function of U . For d >> L , ns is practically independent of U . In order that
the accumulated charge density ns can be tuned by means of U as effectively
as possible, we must have d << L. With U s - UO M 1 V and mz M m , the
corresponding value of L is about 300 A. The insulator layer must therefore
be extremely thin. In the limiting case d << L and Us >> Uo, we have
620 Chapter 7. Semiconductor junctions under non-equilibrium conditions

(7.119)

With U M 5 V , this yields (dn,/dU) M 1013 crn-'V-l. A voltage increase


of 1 V results in about lOI3 electrons being induced into the inversion layer
per cm2.
Electrons in an inversion layer have been the subject of many physical
investigations. Like carriers at a semiconductor heterojunction (see section
6.2), these electrons are confined in a potential well. Just above the bottom
( r i 0.1 e V ) , the width of this well is still smaller than L. Thus, carriers
of such energy are freely mobile only parallel to the well. One has a quasi
2-dimensional electron gas. In it, confinement effects, like the formation of
subbands, can be observed. By means of an applied voltage, the density of
the electron gas can be varied. In this context, the accumulation layer at the
interface between SiOz and Si has been of particular interest. Historically, it
was the 2D gas of this layer in which carrier confinement effects were studied
first, and in which the quantized Hall effect was discovered.

7.4.3 MOSFET
Whether or not the charge density of a semiconductor inversion layer suffices
to cause a resistance change of sufficient magnitude to function as a transis-
tor, depends (among other things) mainly on the specific resistivity of the
semiconductor material in the absence of an applied voltage. This must be
as high as possible. The highest possible resistivity, or the smallest conduc-
tivity, of a semiconductor material is observed when it is in its intrinsic state.
This fact is used to advantage in the most important type of field effect tran-
sistor, the so-called MISFETs (Metal Insulator Semiconductor Field Effect
Transistor). Here, one exploits the fact that in a pn-junction a depletion
region is formed wherein the charge carrier concentrations have intrinsic val-
ues. The p-region is embedded between two n-regions, as shown in Figure
7.22. If one applies a positive voltage between the left n-region (source) and
the right n-region (drain) (Figure 7.22), then the left pn-junction is biased in
the flow direction, and the right in the blocking direction. Between the two
n-regions, one therefore has a p-region which is almost completely depleted
of holes. In the theory developed above, this may be taken into account
formally by adding a positive prevoltage Uv to the voltage U applied to the
insulator-semiconductor junction from outside, i.e. by the replacement

u-+utuv. (7.120)

For U = 0, the semiconductor region between source and drain, the so-called
channel, has high intrinsic resistivity. If a positive voltage U = UG is applied
between the bulk of the p-semiconductor (substrate) and the metal layer on
7.4. Insulator-semiconductorjunction in an external voltage 62 1

a1

I P
Substrat

Figure 7.22: Structure of a MISFET (a). In part (b) the switching scheme of the
MISFET is shown.

..
.
E
n
v)
Y

-0 4 8 12 16 20 24

Figure 7.23: Influence of the gate voltage UG on the I s 0 - versus - U s 0 current-


voltage characteristics for a n-channel MOSFET of enhancement-type. (After Vzilz,
1986.)

top of the insulator (gate), large enough to create an inversion layer, then
the channel becomes a good conductor. The current in the source-drain
circuit can be tuned by the voltage UG of the gate-substrate circuit since the
electron density of the inversion layer depends on UG. The voltage UG is
called gate voltage, and the minimum gate voltage UG necessary to a achieve
inversion is the threshold voltage. If one also includes a working resistor
in the source-drain circuit and compares the power changes in this resistor
with those in the gate-substrate circuit, one finds that the former exceeds the
latter appreciably. In this sense, the MISFET operates as an amplifier, just
like the bipolar transistor. Because of the insulating layer below the gate
electrode, the MISFET has a much larger input resistance than the bipolar
622 Chapter 7. Semiconductor junctions under non-equilibrium conditions

transistor. The output resistance of the MISFET, i.e. that of the source
drain circuit, is small because of the accumulation layer between source and
drain.
If the MISFET is realized using Si as the semiconductor material and
SiO2 as the insulator, one has the Metal Oxide Semiconductor Field Effect
Transistor, abbreviated MOSFET, which is by far the most important MIS-
FET. The basic requirement that the electrons of the inversion layer shall
not be captured by interface states, are met by the MOSFET extremely
well. Just as the base of a bipolar transistor can be made either of p or
n-material, one also has two possibilities in the case of the unipolar MOS-
FET - the charge carriers in the conducting channel can either be electrons,
as has been assumed thus far, or holes. In the first case one speaks of a
n-channel MOSFET, and in the second of p-channel MOSFET. The prevolt-
age U V ,which was introduced above only formally to simulate the intrinsic
state, can really exist for various reasons, e.g., because of positively charged
centers within the oxide arising during its formation. The prevoltage can
take such large values that inversion exists even without an applied gate
voltage. Then the transistor is already in its conducting state at zero gate
voltage. To switch it into its blocking state, one must apply a negative gate
voltage in the case of the n-channel MOSFET, and a positive gate voltage
in the case of the p-channel MOSFET. One says that the transistor oper-
ates in the depletion mode. If the transistor is blocked without an applied
voltage and changes into the conducting state by applying a positive (for
n-channel) or negative (for p-channel) gate voltage, one has the enhance-
ment mode. The current-voltage characteristics of the source-drain circuit
of a n-channel-enhancement-mode MOSFET are shown in Figure 7.23 for
different gate voltages. One recognizes how the sourcedrain current at a
h e d value of U s 0 increases with increasing gate voltage UG. Somewhat
unexpectedly, for a k e d gate voltage UG, the current saturates at higher
source-drain voltages USD. This is a consequence of the fact that the effec-
tive gate voltage in the vicinity of the drain electrode becomes smaller and
smaller as the drain potential grows larger. This creates a pinch off of the
inversion layer, making the current stay constant.
The MOSFET represents the most important electronic component of
digital circuits in microelectronics. Here it is used as an electrically control-
lable switch, meaning that its control function is reduced to two states only,
one with maximum output power, corresponding to a binary l,and another
having minimum output power, corresponding to a binary 0. Today, MOS-
FETs can be made as small that millions can be integrated in one single
Si chip, more than of any other electronic component. It is this ultra-large
scale integration (ULSI) technique which has made modern computer and
communication technologies possible.
623

Appendix A

Group theory for


applications in
semiconductor physics

A.l Definitions and concepts

A.l.l Group definition


A group G is defined as a set of elements g which possesses the following
properties:
1) There exists an assignment instruction which associates an ordered pair
of elements g1 and 92 of the set uniquely with another element which also
is a member of the set. One terms such an assignment instruction a 'mul-
tiplication', and writes g 1 . 92 for the associated element. The totality of
the assignment instructions forms the multiplication table of the group. This
table determines the individual nature of a group.
2) The multiplication obeys the associative law, i.e. for three arbitrary
elements 91, 92, 93 one has the identity

3) The set of elements contains the identity element. This is an element e


defined by the property that its right or left product with any other element
g of the group yields again g. Thus,
624 Appendix A . Group theory for applications in semiconductor physics

4) For each element 4, the set also contains its im-erse element 9-l: such
that
g.g-' =; g -1 - g - e . (A.3)
Onc may show that it suffices to dcmand the existence of only the left-sided
or only the right-sided inverse. This one-side inverse is thcn uniquely deter-
mined and its existcnce automatically leads t o the existence and uniqueness
of the other-siticd iiiverse. T h e other-sidd inverse thusly determined is iden-
tical with the original oneside in?-erse. Analogous dstements hold or the
identity dement discussed above.
The inverse (91 . g2)-' of tl product gl . g:! is given by gz1 . gl1: as one
may easily verify by explicit. mult.iplicat.ionof t h e two products.
The existence of the inverse ensures that t.he products g1 . g and g2 . g
of two elements g1 and g~ with an arbitrary group element g are different,
if the elements g1 and 92 differ from each other. The same is true for the
re-ordered products g .g1 and 9 92.
The number N of elements of a group is callrd its order. Depending on
whether N is finite or infinite, one has finiteor iu,~%ittgrotqs. ' h e dements
of infinite groups may be either discrehe or continuous. In the latter rase
one has a coatinuow group.
The group multiplication which i s associative by definition, need not also
be commutative. In general: g1 . $2 differs from g2 . gl. If the multiplication
is also commutative, the group is call& A h e l i a ~ ~
We now introduce some goup-thcoretical concepts which are n d e d in
this book.

A.1.2 Concepts
A suihject of a group which forms a group by itself with respect to the
multiplication tahle defined by that of the full group, is termed a subgroup.

Subgroups

For finite groups G, the theorem of Lagrange holds, wherein the order 1L"
of a subgroup G I is a divisor ol thp order N of the group. Thc proof of

this theorem will be briefly sketched because it also provides some insight
in other group properties. Let be ,g! = E,g$,. . , g h , the elements of a
subgroup G" of G We denote an arbitrary element of G not contained in
G' by gz. 'I'he products 92 gi, g 2 . gh,. . . , g2 . gh, of g:! with dl elements
of the subgroup G' form a 5et M,<, the elements of which are different from
each other as well as from the dementb of G'. For the sake of uniformity
of notation, w e set C' M,<. If the two sets M1< and M; together do
not cover trhed i r e group C , WP choose an element 93 of G not contained in
AT,< and M,<. We proceed in the same way with this element as we &d with
A.1. Definitions and concepts 625

92. The thus emerging set M3< contains N' mutually different elements, just
like M,< and M 2 , and none of these elements occurs in M1< and h42. If the
entirety of the sets M:, and M Z still do not yet cover the full group
G, we choose yet another element 94, and so on. After a finite number N f f
of steps a complete division of the group G into N" distinct subsets with N'
elements in each will be achieved. Therefore, one has N = N' ' N" which
proves the theorem.
Lagrange's theorem allows us to draw several conclusions. For groups
of prime order it means that only trivial subgroups can exist, namely the
identity element, and the group itself. This statement may also be reversed
- the order of a group which has no subgroups apart from the trivial ones,
must be a prime number.

Cyclic subgroups

Cyclic subgroups consist of integer powers ,'g k = 0,1,. . . ,K - 1, of an


element g where K is a positive integer. For finite groups the value of K is
finite and g K coincides with the identity element. If the cyclic subgroup with
respect to a certain element is the entire group, then the same necessarily
holds also for the cyclic subgroups with respect to all other elements. The
group itself is then said to be cyclic. Cyclic groups are also Abelian. A
group whose order is a prime number must necessarily be a cyclic group - if
this were not so, the group would have subgroups, cyclic ones, which is not
allowed in this case.

Normal divisor and factor group


The sets M;, i = 1 , 2 , . . . , N" defined above are called left cosets of G with
respect to the subgroup G' for the elements gi. Analogously, one defines
right cosets M,' of G with respect to the subgroup G' for the element gi,
namely as sets of elements g: ' g i , g a . g i , . . . , g h , . g i . If all the right cosets with
respect to a particular subgroup G' are identical with all the left cosets with
respect to this subgroup, then the subgroup G' is called a normal divisor.
One can readily see that such a subgroup consists only of classes of conjugate
elements of G (for the definition of classes see below).
We can go a step further and understand the cosets with respect to
the normal divisor G' as group elements and define the product of two such
elements as the set of the products of all their elements. Since these products
are again cosets, the multiplication thusly defined constitutes a group. It is
called factor group. Normal divisors and factor groups are helpful concepts
for the characterization of the internal structure of groups.

Direct product

We consider two groups G and G' of, respectively, orders N and N', and
626 Appendix A . Group theory for applications in semiconductor physics

examine the products between all elements g k of G and all elements gh, of
G'. The whole set of these N . N' products g g i , = g k . gLi forms a group G"
of order N . N'. This statement may be easily verified by showing that the
product of two of the elements of GI' is again an element of G". The group
GI' is called the direct product of the two groups G and G I , for which one
writes G" = G x G'.

Conjugate elements and classes


In regard to representation of groups by matrices, which we will consider
later, the division into classes of conjugate elements is of great importance.
An element g 1 is said to be conjugate to an element 92, if the group contains
an element g such that g 2 = g . g 1 . 9 - l . From this definition it follows
immediately that if g 1 is conjugate to 92, then 92 is conjugate to 91, and g l l
is conjugate to g T 1 . Moreover, it holds that with g 1 conjugate to g 2 and 92
to 93, then g 1 is also conjugate to 93.
Now we consider a particular element g 1 and form all elements g . g 1 . g - l
conjugate to it. This may be done by allowing g in the latter expression to
run through the entire group. Thereby, a certain set of elements K1 emerges.
Consider an arbitrary element of this set and again form all elements conju-
gate to it. The set of elements thusly obtained coincides with the previously
obtained set K1. One calls K1 a class of conjugate elements or simply a
class. If one applies the same procedure to an element g 2 of the group not
contained in K1, one obtains a set K2 which has no common elements with
K1. It forms another class of conjugate elements. If one continues to apply
this procedure until no element remains which does not already lie in one of
the prior classes, one achieves a complete division of the group into distinct
classes of conjugate elements. One may say, therefore, that each group con-
sists entirely of classes of conjugate elements. Each class is defined by one of
its elements, the remaining elements follow by conjugating the defining ele-
ment. The identity element forms a class by itself in all groups. In Abelian
groups each element forms its class.
Homomorphism and isomorphism
We may associate the elements g of the group G uniquely with the elements
f of another group F , as expressed by the mapping relation

f = f( 9 ) . (A.4)
Such an association, or mapping, is termed homomorphic if it obeys the
condition
f ( g 1 ' 92) = f (91) . f (92) 7 (A.5)
where g1 and 92 are two elements of G. If the mapping is not only unique but
is also bi-unique, i.e. if not only each element of G corresponds to exactly one
A.2. Rigid displacements 627

element of F , but also, conversely, each element of F corresponds to exactly


one element of G, then the mapping is isomorphic. Two groups which are
related to each other by an isomorphic mapping, are called isomorphic. Since
groups are completely characterized by their multiplication tables, and since
the latter are necessarily the same for isomorphic groups, one may say that
two isomorphic groups are essentially two different realizations of one and
the same group. They differ only in the meaning of the elements.
For the description of microscopic properties of crystals, three kinds of
groups are of particular interest the translation groups, the point groups
~

and the space groups. The elements g of these groups are in all three cases
rigid displacements of certain point sets, which transform these sets into
themselves. As we will see, the point sets for the translation groups are
infinite lattices, those for the point groups are finite spatial bodies, and
those for the space groups are crystals, i.e. several infinite lattices of the
same Bravais type put into each other. The translation-, point- and space
groups are summed up under the term displacement groups.

A.2 Rigid displacements

A.2.1 Definition
Formally one may define rigid displacements as transformations g of 3-
dimensional space which shift its points x by vectors a(x) in such a way,
that the distance I x1 - x2 I between two arbitrary points ~ 1 ~ does
x 2 not
change. Thus, we have

or
2
2(x1 - x2) Ia(x1) - a(xz)l+ I 4x1) - a(x2) I = 0. (A4
The multiplication dot in (A7) and (A.8) signifies the internal or scalar
product. In order to satisfy equation (A.8), the displacement field a(x)
cannot be of higher order in x than of the first, i.e. a(x) must have the
general form

a(x) = px +a (A.9)
where /3 is a linear operator in x-space and a a constant vector. From (A.6)
it follows that
628 Appendix A. Group theory for applications in semiconductor physics

gx = a x + a, (A.lO)
+
with a = 1 0. The thusly defined operator cr is also linear. It is subject
to certain constraints due to the rigidity condition (A.7). If one enters this
equation with expression (A.lO) for gx,then one gets, first of all, (XI- xg).
( X I - xg) = a(x1 - xg) . a(x1 - x2). Using the adjoint operator a+ of a,
the last relation may be put into the form

(XI - xg) ' (XI - x2) = a+ ' a(x1 - x2) . (XI - x2). (A.ll)
The multiplication 'dot' between the two operators symbolizes the consecu-
tive application of the operations. Since (XI - xp) represents an arbitrary
vector, relation ( A . l l ) implies that

a + . a = 1. (A.12)
By this identity, a+ is seen to be the inverse operator a-l of a , whence

+
, = ,-I. (A.13)
Linear operators which act in a real vector space (as we assume here) and
satisfy the relations (A.12) or (A.13), are called orthogonal. The transfor-
mation they describe are orthogonal transformations.

A.2.2 Translations
For p = 0 or a = 1, g reduces to a rigid displacement of all points through
the same vector a, i.e. to a translation. We use the symbol t, for it, so that

tax = x + a. (A.14)
With ta, one can rewrite equation (A.lO) for gx in the form

(A.15)
whence
(A.16)
which is stated as theorem below.

Theorem

Any rigid displacement g m a y be generated by a n orthogonal transformation


a followed by a translation t,.
A.2. Rigid displacements 629

This theorem remains valid if one reverses the sequence of the orthogonal
transformation and the translation, but one must then use another transla-
tion in order to obtain the same displacement g as before, specifically

t, ' a = a!.ta-la. (A.17)

A.2.3 Orthogonal transformations


Orthogonal transformations a have a number of important properties, which
we will now explore in detail. For this purpose we introduce a Cartesian
coordinate system with orthogonal unit vectors e l , e 2 , e 3 . The effect of a
on the basis vectors can be described by means of a matrix A ( a ) having real
elements given by

aik(a) = ei.a e k . (A. 18)


The transformation a itself may be expressed in terms of the dyadic struc-
tures e i e k composed of two vectors e i and e k as follows:

a =C C a i k ( a ) e z e k . (A. 19)
i k

The application of a as given in equation (A.19) on a vector v is defined in


terms of the formation of the scalar product of e k with v, i.e.

av = c
i
Cai&)eiek
k
' v. (A.20)

It follows that

aei =Caki(a)ek=Ca$(a)ek. (A.21)


k k

where a:(.) signify the elements of the matrix [A(a)IT,the transpose of


A ( a ) . Using of A ( a ) , we may also describe the effect of the transformation
a on a point

x =Cxiei, (A.22)
1

where
xi = e i . x (A.23)
are the Cartesian coordinates of x. We obtain

(A.24)
630 Appendix A. Group theory for applications in semiconductor physics

By this relation one may understand the coordinates to be the transformed


quantities instead of the basis vectors, as an alternative interpretation. The
corresponding transformation equation reads

(A.25)

This is to say that the coordinates themselves transform with the matrix
A ( a ) , in contrast to the basis vectors whose transformation is governed by
the transposed matrix [ A ( a ) I T .
The product 0 1 . a2 of two orthogonal transformations a1 and ag, un-
derstood in the sense of sequential application, is represented by a matrix
A ( a 1 . ag) which one can easily determine with the help of relation (A.19),
obtaining

i.e. the matrix of the product transformation is represented by the product


of the matrices of the two individual transformations.
Furthermore, other properties of the transformation operators a are car-
ried over to the corresponding matrices A ( a ) . In particular, it follows from
equation (A.13) that [ A ( a ) ] + = [ A ( a ) ] - l holds. Since the adjoint matrix
[A(a)]+ of a real matrix A ( & ) equals the transposed matrix [ A ( a ) l T ,one
also has

[ A ( a ) l T= [ A ( a ) ] - l . (A.27)
Real matrices with this property are called orthogonaL Thus orthogonal
transformations are represented by orthogonal matrices with respect to or-
thonormalized basis sets.
Relations which follow immediately from (A.27) are A . AT = 1 and
A T . A = 1, which may be written in terms of matrix elements as

This means that orthogonal matrices have the property that different rows
are orthogonal to each other, and so are different columns.
The product of two orthogonal matrices is again an orthogonal matrix.
Combining this result with equation (A.26), we arrive at the conclusion
that the product of two orthogonal transformations is also an orthogonal
transformation. The implicit assumption behind this result is that both
transformations have a common point which remains unchanged (the point
x = 0 here).
A.2. Rigid displacements 631

Figure A.l: Definition of the


Euler angles.
t$

A.2.4 Geometrical interpretation


Orthogonal transformations may be interpreted geometrically. In order to
develop such an interpretation one needs some results concerning the eigen-
values and eigenvectors of orthogonal matrices. As is known from linear
algebra, these matrices have complex eigenvalues with magnitude 1. The
corresponding eigenvectors form a basis in 3-dimensional complex vector
space. In this space the matrices can be transformed in such a way that
they become diagonal (in the real 3-dimensional vector space this is not pos-
sible). For each eigenvalue X the corresponding complex conjugate A* is also
an eigenvalue, so there must be at least one real eigenvalue of magnitude 1.
To this real eigenvalue corresponds a real eigenvector e. The real eigenvalue
of unit magnitude can only be either +1 or -1. The determinant D e t [ A ]of
A is +1 or -1, respectively. In the case of D e t [ A ]= +1 one has

ae = e. (A.29)
Since only the direction, but not the length, of e is determined by this
relation, equation (A.29) means that all points given by xo = <e with <
as arbitrary real number which lie on a straight line in the direction of e
through the origin, are transformed into themselves. One says that they are
fixed points of the transformation a. The orthogonal transformation (Y with
Det[A]= +1 therefore corresponds to a rotation about the axis defined by
xo. A special rotation is the identity transformation a = E .
In the case D e t [ A ]= -1 one has
632 Appendix A. Group theory for applications in semiconductor physics

ae = -e. (A.30)
The transformation a therefore describes a rotation about the axis xo = e
followed by a reflection in the plane normal to the rotation axis through the
origin. It is called an improper rotation or a rotation-reflection . Among the
improper rotations are also pure reflections in the plane normal t o the axis,
and inversion with respect to the origin. In the first case the rotation angle
is 27r, and in the second case T. Since an improper rotation is constituted
of the consecutive application of a rotation and a reflection, all orthogonal
transformations may be traced back to pure rotations and pure reflections.
The orthonormality relations of (A.28) mean altogether six independent
equations. The nine elements of an orthogonal matrix may thus be thought
of as functions of three independent variables. The choice of these vari-
ables is not unique. The most common is the one by Euler, who uses three
angles $,8 and p. Two of these angles, $ and 8, determine the direction
(sin 8 sin $, - s i n 0 cos $, cos 0 ) in which the z-axis is turned, and cp is the rota-
tion angle about the transformed z-axis (see Figure A.l). The corresponding
transformation matrix reads

A= ( cos $ cosp

sin$ cosp
sin 6 sin p
- cos 6 sin$sinp
+ ws6 cos $ sin9
- cos$ sin p - cos 6 sin $ cos p sin6 sin$
-sin $ sin p + cos6 cos $cos
sin 6 cos p
p - sin 6 cos $
cos 6

Using this matrix, a reflection in a plane normal to the direction (sin8 sin$,
- sin 8 cos $, cos 0) may be traced back to a reflection in a plane normal to
the z-axis, which is easily described. The matrix A of the reflection in the
plane normal to the above cited direction thus reads

cos 28 sin $ + cos $ sin 2$ sin 0 - sin 28 sin $


cos 28 cos $ + sin $ sin 28 cos $
- sin 28 sin $ sin 28 cos $ - cos 28
(A.32)

A.2.5 Screw rotations and glide reflections


The geometrical interpretation of the orthogonal transformations makes it
possible to provide concrete illustrations of the theorem on the decomposi-
tion of rigid displacements into translations and orthogonal transformations.
One can show, specifically, that a rotation about a particular axis followed
A.2. Rigid displacements 633

b) I

Figure A.2: a) Rotation followed by a Figure A.3: a) Rotation followed by


translation normal to the rotation axis, a translation parallel to the axis (screw
b) Reflection followed by a translation rotation), b) Reflection followed by a
normal to the mirror plane. translation parallel to the mirror plane
(glide reflection).

by a translation normal to the rotation axis, can be represented as a pure


rotation around an axis parallel to the former, but shifted with respect to it
in the normal plane. A reflection with respect to a particular plane followed
by a translation normal to the plane can be represented as a reflection in a
plane parallel to the former, but translated with respect to it in the normal
direction. Using Figures A.2 and A.3 one can readily convince oneself of the
validity of these assertions. A formal proof may be given as follows.
We consider, first of all, rotations followed by translations normal to the
rotation axis. In this case, the transformation g is of the general form (A.10),
where the hed-points xo of cy are solutions of the equation

ax0 = XO, (A.33)


634 Appendix A . Group theory for applications in semiconductor physics

with
x o a = 0. (A.34)
The identity element Q = E must be excluded here because no rotation
axis is defined for it. What has to be proved is that, under the conditions of
equation (A.33), the transformation g itself has hed-points xb which form
a straight line, i.e. that the equation

cvxb+a=xb (A.35)
has a 1-dimensional solution manifold. To verify this, we express xb in the
form

+
xo = (2x0 ra + sa-la (A.36)
with q , r, s as coefficients which still have to be determined. Since the three
vectors xo, a and a-la, are linearly independent, such a representation is
always possible expect for the special case aa = -a in which the rotation
angle p of Q about the axis xo is equal to T. The latter case has to be
considered separately. Employing equation (A.36) for xo jointly with equa-
tion (A.35), we obtain a vectorial relation for q , r and s which is satisfied by
arbitrary values of q . Consequently, the result is an inhomogeneous vector
equation in the plane perpendicular to xo for r and s only. Considering the
components of this equation with respect to a and cra, we find a system
of two inhomogeneous equations, whose determinant equals laI2 2 sin2 cp. It
never vanishes since the two cases cp = 0 and cp = T have been excluded, and
thus there is a unique solution for r and s. In the excluded case cp = n one
may take s = 0, and then the above procedure yields a unique solution for r .
With this, our assertion is proved for a rotation with consecutive translation
normal to the rotation axis. In the case of a reflection accompanied by a
translation normal to the mirror plane, the proof is analogous.
There remain rotations and reflections followed by translations parallel
to, respectively, the rotation axis and the reflection plane. That these cannot
be replaced by pure rotations or reflections is obvious. The rotation axis
and the reflection plane are unaffected by the translation. Therefore a pure
rotation would have to be one about the same axis, and a pure reflection one
with respect to the same plane. But these have to be followed by translations
in order to get the complete transformations. Therefore pure rotations and
reflections do not exist in this case.
In summary, this means that rigid displacements can be traced back
to rotations and reflections only if translations which are directed parallel
to the rotation axis or mirror plane are not involved (see Figures A.2 and
A.3). Rotations followed by translations parallel to the axis are called screw
rotations, and reflections followed by translations parallel to the mirror plane
A.3. 'Ikanslation,point and space groups 635

are glide reflections. These results allow one to express the above stated
theorem in the following, more specific form:

Theorem 2

A n y rigid displacement is either a screw rotation ore a glide reflection, o r a


product of the two.

A.3 Translation, point and space groups

A.3.1 Lattice translation groups


Translations which transform crystal lattices into themselves only may have
displacement vectors a equal to lattice vectors R. We will show that the set
of all lattice translations t R forms a group. The product t R 1 . t R , of two
translations tR1 and t R z is defined as the translation which results from the
consecutive application of the two individual translations. One may express
this by writing

which yields
tR1 ' tRz = t R ~ + R ~ (A.38)

The latter relation means that the product of two lattice translations defined
in equation (A.37) is again a lattice translation. This set therefore forms a
group if points 2), 3) and 4) of the general group definition are also fulilled.
That this is so may be demonstrated easily. Using relation (A.4) one can
readily show that the product (A.38) obeys the associative law (point Z),
that an identity element exists, namely the translation through the lattice
vector 0 (point (3)), and that the inverse of each element t R belongs to the
group, namely the translation t-R through the lattice vector -R (point 4).
Each lattice translation can be expressed as a product of translations
through three primitive lattice vectors al,az, ag, so that

(A.39)

Since the sequence in which the primitive translations are executed plays
no role, all other lattice translations also commute with each other, i.e. the
translation group is Abelian.
636 Appendix A . Group theory for applications in semiconductor physics

A.3.2 Point groups

Definition

Point groups are sets P of orthogonal transformations a , namely of rotations,


reflections, rotation-reflections , rotation-inversion s and the inversion itself,
which transform geometric bodies having planar boundaries, i.e. prisms,
tetrahedrons, cubes or other polyhedrons, into themselves. One may thus
define point groups as the symmetry groups of these bodies. In this context
their elements are called point s y m m e t r y elements or point s y m m e t r y oper-
ations. That the set of all point symmetry operations of a geometric body
actually forms a group, may be seen as follows. Firstly, as has already been
demonstrated above, the product a1 . a2 of two orthogonal transformations
a1 and a2,understood as their consecutive application, is itself again an or-
thogonal transformation. The condition under which this statement holds,
namely the existence of common fixed-points, can be satisfied in the case at
hand. One has to keep in mind, however, that the various symmetry rotation
axes and mirror planes of the geometric body have to be chosen to have at
least one point in common. Secondly it is also clear that the transformation
a1 . a2 maps a body into itself if this is true for a1 and a2 separately. Thus
the product also belongs to the set P of all point symmetry operations as
well. The validity of the associative law (a1.a2).ag= al.(aq.ag) is evident,
likewise the existence of a identity element and of the inverses. This means
that, with respect to the above defined product, P does in fact form a group.
Since rotations about two different axes and reflections at different planes
do not, in general, commute the point groups are in general non-Abelian.
The number n of the different rotations about a particular axis is finite in
all point groups. From this it follows that the rotation angles may only differ
by multiples of (an/.). Moreover, only a finite number of different rotation
axes and mirror planes occur. Point groups are therefore finite groups. This
follows from the restriction to bodies with planar boundaries. If we would
had also allowed for the sphere as geometrical body, then the symmetry
operations would have included rotations about arbitrary axes and through
arbitrary angles, likewise for reflections at arbitrary planes. Then one would
no longer have, however, a point group, but the full orthogonal group. If
one removes from the latter group all elements which involve a reflection or
inversion and retains only pure rotations the remainder is the full rotation
group. Both groups, the full orthogonal group and the rotation group, are
infinite. The point groups may be understood as finite subgroups of the full
orthogonal group.
Beside geometric bodies having planar boundaries, there are yet other
point sets which are transformed into themselves by point groups. In the
A.3. Tkanslation, point and space groups 637

main text we have shown that such sets include infinite point lattices. In
this case, however, only 7 special point groups are admissible as symmetry
groups. As far as ideal crystals are concerned, these are neither geometric
bodies nor infinite lattices. Therefore, the symmetry groups of crystals, the
so-called space groups, are not necessarily constituted of point groups. This
is not so for the symmetry groups of equivalent crystal directions, which
were also considered in the main text. These are point groups, but again
only special point groups are possible, altogether 32. One calls them point
groups of equivalent crystal directions or, in short point groups of directions.

Description of point groups

Generating symmetry elements

The lattice symmetry translations can be set up as products of translations


through the three primitive lattice vectors. It turns out that likewise also all
point symmetry groups can be generated from a small number of elements
by forming products of them. The set of generating elements in the case
of point groups is, however, larger than for translation groups. In neither
of the two cases is the choice of generating elements unique. Here we con-
sider the transformations which are commonly used for the construction of
point groups. Their actions are illustrated in Figure A.4. A list of these
transformations follows:
a) Rotation C, by an integer fraction 2 n l n of 2 n about a vertical axis. As
vertically one refers to the axes of highest symmetry. Thus it depends on
the point group which axis is named vertical. On each axis one defines a
direction to which the axis points, so that rotations by positive and negative
angles can be distinguished. Positive angles correspond to a right handed
screw sense of the rotation. The whole set of rotations by 2 ~ / about n the
axis is called an n-fold rotation axis or a rotation axis of multiplicity TL The
symbol for such an axis in the Schonflies notation is C,, just as for the
single rotation through 2nln as indicated above. Commonly, one tolerates
this ambiguity of notation and adds, if necessary, whether the axis is meant
or the rotation. The international notation for an n-fold axis is n.
b) The rotation U:! by n about an axis which forms a right angle with a
vertical axis, i.e. about a horizontal axis. In the international notation
scheme a horizontal axis is abbreviated by a 2 which follows the symbol n
for the n-fold vertical axis to which it belongs, as in n2.
c) Reflections nV or Od in a plane which contains a vertical axis. This plane
is called a vertical plane or a diagonal plane. In the international scheme
one uses an m for such a plane, following the n for the vertical axis which
it contains, as in nm.
638 Appendix A. Group theory for applications in semiconductor physics

ml
cn .t
-___---
"2

b)
I
I

el

Figure A.4: Generating symmetry elements of the 32 point groups of crystals.


A.3. ZhnsJation, point and space groups 639

?:'
Sni
I I

f)

Figure A.4: Generating symmetry elements of the 32 point groups of crystals


(continuation).

d) A reflection u h in a plane orthogonal to a vertical axis, i.e. in a horizontal


plane. The international notation is an 'm' under the 'n' for the vertical
axis, as in
e) A rotation by 2 ~ / about
n a vertical axis followed by a reflection at the
horizontal plane ffh containing the axis, i.e. a rotation-reflection . The
Schonflies symbol for this element is Sn. By definition we have

sn=Uh' cn=cn ' Oh' . (A.40)


The two factors in this product commute because the reflection U h does not
change the directionof the rotation axis and the absolute value of the rotation
angle, if it is applied before instead of after the rotation. The direction to
which the rotation axis points is reversed, of course, however, with this, the
sign of the rotation angle is also changed, so that in both cases we have the
same rotation.
The symbol Sn is also used to denote an n-fold rotation-reflection axzs,
which means the whole set of powers of 27r/n.
f ) A rotation by 2 ~ / nabout a vertical axis followed by an inversion I at
640 Appendix A . Group theory for applications in semiconductor physics

Figure A.5: Principle of the


stereographic projection.

Figure A.6: Graphical sym-


bols for symmetry elements of
point groups. Shown in the left
column are 2-, 3-, 4- and 6- A A
fold rotation axes, in the mid-
dle column 2-, 3-, 4 and 6-fold
rotation-reflection axes, and in
the right column vertical and di-
agonal mirror planes. * @
the fked-point (see Figure A.4g), or an rotation-inversion The Schonflies
symbol for it is S,i. By definition we have

S,i = I . c, = c,. I. (A.41)

The two factors commute for the same reason as in the case of rotation-
reflections considered above. Rotation-inversion axes, which are understood
as the set of all powers of S,i, are used in the international system. The
notation for an n-fold rotation-inversion axes is E.

Stereograms

Before investigating the various point groups, we will provide a procedure


for the 2-dimensional illustration of these groups, namely the method of
stereograms. This method relies on the fact that the elements of point groups
are orthogonal transformations which transform a point on the surface of a
A.3. Banslation, point and space groups 64 1

sphere around the common fixed-point into another point on the surface
of the sphere. The elements of a point group are thus put into one-to-
one correspondence with points on the surface of a sphere. At best, one
imagines the sphere as a globe. To achieve a planar representation) one
utilizes, as in cartography, the stereographic projection. This is illustrated
in Figure A.5 with the south pole as the reference point. In this case the
northern hemisphere is projected onto the region within the equator, and
the southern hemisphere onto the region outside the equator. In order that
only interior points need to be considered) one takes the north pole as the
reference point for projection of the southern hemisphere. Points of the
northern hemisphere are shown as full circles in the projection, and points
of the southern are blank circles (see Figure A.5). It can be shown that the
stereographic projection preserves circles and angles, i.e. circles (including
straight lines as circles of infinite radius) are projected into circles, and the
angles between intersecting tangents of circular segments also remain the
same in the projection. As for the elements of the point group, one can
also represent rotation axes and mirror planes by means of stereographic
projection. Figure A.6 displays the corresponding graphic symbols. A n-
fold rotation axis, for example, is depicted as a small regular n-angle placed
at the projection of the intersection of the axis with the northern hemisphere.
A mirror plane is illustrated by the projection of the arc on which the plane
intersects the northern hemisphere. That the equatorial plane is a mirror
plane may be recognized by the fact that for each full circle one has a blank
circle at the same place. The procedure described here allows a 2-dimensional
illustration of the point groups. The result of the projection is called a
stereogram.
The stereograms may also be used to obtain the product of two group
elements. As an exercise, the reader should try to verify, in this way, the
following helpful relations between generating elements:

u2 . U v = U v .u2 = U h ,
u2. U h = U h . u2 = U V ,

where U2 is the horizontal axis given by the intersection of the two planes
uv and Uh. Analogous relations hold with U d replaced by uv. Furthermore,
one has
Uh.c 2 = c 2 Uh = I ,
' (A.43)
I .Uh Oh ' I =c 2 , c2 ' I = I . c2 = U h ,
I . uv = U V 'I = u2 , u2. I = I . u2 uv.
Here U2 is understood to be orthogonal to uv.
6 42 Appendix A . Group theory for applications in semiconductor physics

Figure A.7: Stereogram of the


point group C3 (3).

General rules for the division of point groups in classes

The division of point groups in classes of conjugate elements is facilitated


by some general rules which follow immediately from the definition of the
conjugation given in section 1. These rules read as follows:
1) Rotations through the same angle with respect to symmetrically equiva-
lent rotation axes are mutually conjugate.
2) If the group contains a symmetry operation which reverses the direction
of a rotation axis, as with a 2-fold axis normal to the rotation axis, or a
symmetry operation which changes the sign of the rotation angle, as with a
mirror plane parallel to the rotation axis, then the axis is called two-sided
Each rotation around a two-sided axis is conjugate to its inverse.
3) Reflections at symmetrically equivalent planes are conjugate with respect
to each other.
Depending on whether a point group contains only rotations or, in addition,
reflections, rotation-reflections , rotation-inversion s or inversion, one speaks
of point groups of the first kand or point groups of the second kind. The point
groups of the first kind may also be understood as finite subgroups of the
full rotation group. Below the full set of these groups will be described.

Point groups of the first kind

cn (4
The simplest point group of the first kind consists of the n powers CA, C$-
. . . , Cg of the rotation C, around a vertical axis. These are rotations
through, respectively, ( 2 ~ / n (2n/n)
), . 2,. . . , (21r/n) . n = 2 ~ The
. rotation
C c is the unit element E. Using the terminology introduced above one may
A.3. Translation, point and space groups 643

Table A.1: Multiplication table of the point group C3.

U U

briefly say that the group consists of an n-fold rotation axis. Thus the nota-
tion for the group is C, in the Schonflies system, and n in the international
system. The multiplication of two rotations Ck and C i leads to

c,. c i = c y . (A.44)
If k + j is larger than n, one may replace C;+j by C,k+j-.. The product thus
belongs to the set of elements CA, C:, . . .. Table A . l shows the complete
multiplication table of the group CJ. The product in its kth line and Zth
column is the result of the multiplication of the lcth element as left factor
and the lth as right. In the group C, the multiplication order plays no role,
because this group is cyclic and thus also Abelian. Each element forms a
class of conjugate elements by itself. The stereogram of the point group C3
is shown in Figure A.7.
Starting from the point group C , one may set up other point groups of
the first kind by complementing the n-fold rotation axis with new generating
elements, along with the products of these elements among themselves and
with the elements of the original group.

D, (n2, n22)
The group C, is first of all complemented by a horizontal rotation U2. This
may also be expressed by saying that a horizontal 2-fold rotation axis is
added, for the identity element is already contained in C,. However, in this
way no group is yet obtained. The multiplication of the powers C$-, k =
1 , 2 , .. . , n, with U2 produces n - 1 new elements, namely U2k = Ck- . U2
(here Uzl is merely another notation for U2). The set of these elements is
identical to the set of elements U2 . Ck- arising by multiplication with U2
from the left, but the numbering is different. The n elements U2k correspond
to n different 2-fold rotation axes. Each pair of adjacent axes forms an an
angle x/n. For odd n all n rotation axes U2k are symmetrically equivalent to
each other, while for even n they fall into two distinct classes of inequivalent
axes, namely U21, U23, .., U2,-1 for odd 1,and U2, U22, .., U2n-2 for even
644 Appendix A. Group theory for applications in semiconductor physics

a) b)
Figure A.8: Stereograms of the point groups D2(222) (a) and D3(32) (b).

Table A.2: Multiplication table of the point group D3

k . One must add all n rotations u 2 k to the group C, in order to get a


new group. This group then contains 2n elements. Their designation in the
international system is n2 if n is odd, and n22, if n is even. The 2 following
the odd n stands for the n symmetrically equivalent horizontal rotation
axes, and the 22 following the even n stands for the two symmetrically
inequivalent classes of axes. In the Schonflies system the designation of the
group is D , in both cases.
One may interpret the group D , as the symmetry group of all rotations
of a straight prisms through a regular n-angle (the full symmetry group
of this body still contains reflections and is therefore of the second kind).
Figure A.8b shows the stereogram of the group D3, and Table A.2 shows
the multiplication table of D3. The latter may easily be checked using the
stereogram in Figure A.8b. The group D2 with three mutually orthogonal
2-fold axes, called also V , is shown in Figure A.8a.
A.3. ?f-anslation,point and space groups 645

Figure A.9: Rotation symmetry of a tetrahedron (left) and stereogram of the point
group T (23)(right).

We still have to specify the division into classes of conjugate elements for the
point group D,. Applying the first rule for class division derived above one
finds that the n 2-fold rotations form one class for odd n , and two for even
n. The second rule implies that a rotation Ck and its inverse CLk belong
to the same class. These form ( n - 1)/2 classes if n is odd, and n/2 if n
is even. In addition one has the identity element E which forms a class by
itself. Altogether, this yields ( n - 1)/2 + 2 classes for odd n, and n / 2+ 3
classes for even n. The group Dz has the 4 classes E , C2, U21 and U22, i.e.
each element forms a class by itself. The group D3 has 3 classes, namely E ,
the three 2-fold horizontal rotations (U21,U22 , U231, and the %fold vertical
rotations (C3, C:} (here and below elements of the same class are shown in
curly brackets).

T (23)
Another point group of the first kind is the group of all rotations which
transforms a tetrahedron into itself (the full symmetry group of a tetrahedron
also contains rotation-reflections , and is therefore of the second kind). The
rotation axes of a tetrahedron are the four %fold vertical axes C3 through
a corner and the center of the opposite face, and the three vertical axes C2
through the centers of opposite edges (see left-hand part of Figure A.9). We
will not name the axes differently. The international symbol of the group
reads '23', and the Schonflies symbol T . The four %fold axes yield a total
of 8 different rotations, 4 of the form C3, and 4 of the form C:. The three
2-fold axes yield a total of 3 elements of the form Cz. Along with the identity
646 Appendix A . Group theory for applications in semiconductorphysics

Figure A. 10: Stereogram of the


point group 0 (432).

element, the group T therefore consists of 12 elements. That the set of these
elements does in fact form a group is most easily verified by means of the
stereogram in Figure A.9. Each of the following sets of elements forms a
class: the 4 rotations C3, the 4 rotations Ci, the 3 rotations C2, and E.
There are, altogether, 4 classes.

0 (432)

The group of all rotations which transform a cube into itself is also a point
group of the first kind (the full symmetry group of a cube is of the second
kind). The group consists of the three 4-fold rotation axes C4 through the
centers of two opposite cube faces, the four %fold vertical axes C3 formed
by the diagonals of the cube, and the six 2-fold axes C2 through the centers
of opposite cube edges. Altogether, this comprises 24 symmetry operations.
They are illustrated in the stereogram of the group in Figure A.lO. The
international symbol of the group is 432, and the Schonflies symbol is 0.
The classes of 0 are the following: all 6 rotations C4 and C: (for the axes
C4 are two-sided), all 3 rotations Cz, all 8 rotations C3 and C i (the C3-axes
are likewise two-sided), and all 6 rotations C2 and E . One therefore has 5
classes. We abbreviate them in the form E , 6C4, 3C2, 8C3, and 6C2, the
symbol for each class being a representative element with the number of
elements in the class positioned in front.
Beside those already mentioned, there is only one other point group of
the first kind, namely the rotation symmetry group Y of a pentagondodec-
ahedron. This refers to a body with twelve faces, each of which forms a
regular pentagon. The group Y , which is also known as the icosahedral
group, contains &fold rotation axes which cannot occur in the symmetry
groups of crystals. Therefore we will not discuss the group Y here further.
A.3. fianslation, point and space groups 647

Point groups of t h e second kind

sn (z, ii), n even

The simplest point group of the second kind consists of one n-fold rotation-
reflection axis s, of even order n. In the Schonflies system this group is
denoted by S,. It contains rotation-reflections S," with k = 1,2, . . . , n.
Since for even n the relation S," = (Uh . C,)" = E holds, there are no other
elements. For odd n one has S," = Oh, which means that the set of the n
powers Sk, with k = 1,2,. . . ,n does not form a group. Thus, we exclude
odd n.
The group S, is evidently cyclic. Each of its elements forms it own class,
for a total of n classes. The even powers Sk describe rotations through
angles ( 2 ~ k / n )= (27rg/?). Thus the n-fold rotation-reflection axis also
automatically generates a (n/2)-fold ordinary rotation axis. This also means
that the group C,/2 is a subgroup of Sn. For even n not divisible by 4, the
group S , contains the inversion I , namely the element S,"12= (Uh.Cn)n'2 =
a h . C2 = I . The group S 2 consists only of I and the identity element E.
One also denotes S 2 as Ci. Each group S, with even n, not divisible by 4,
can be understood as the direct product C,l2 x Ci of the two groups C,/z
and Ci.
In the international notation system one employs, instead of rotation-
reflections s,, the rotation-inversion s S,i as generating elements. In the
case of even n, the powers Sii with k = 1, 2, . . . , n already form a
group, because then Szi = E holds. For odd n one needs the 2n powers
Ski, k = 1, 2, . . . , 2n, to get a group. In both cases the international
symbol reads 6. If n is odd, the group ii is equal to S2n For even n one
must distinguish between those n which are divisible by 4 and those which
are not. In the former case, we have ii = S,, i.e 4 = S4, 8 = SB etc. In the
latter case the group fi can be generated by a (n/2)-fold rotation axis and
a vertical mirror plane q. The group 2 consists of only this mirror plane,
which implies that it contains only the two elements E and f f h . It is denoted
by C,. Generally, the group + i be written as the direct product C , / 2 x C ,
can
of C,p and C , if n is even and not divisible by 4.
Applying the above results to our original groups S, with even n, we
recognize that they can all be traced back to 6' groups. One has S2 =
T, S 4 = 1,s6 = 3, s8 = 8, Slo = 5, S12 = Tz etc. Some of these point groups
are illustrated by their stereograms in Figure A.ll.

The group C, gives rise to a new group if one adds the reflection O h at the
plane normal to the rotation axis. It contains besides the n elements C,"
648 Appendix A . Group theory for applications in semiconductor physics

@
Figure A.11: Stereograms of the point groups Sz(T) = Ci, S4 (4), s6 (3), and
2 = c,.

with k = 1 , 2 , . . . , n ,also the n elements ffh . ck = ck . ffh. Its order


is thus 2n. The international symbol reads E, and the Schonflies symbol
Cnh. The groups Cn and C , are subgroups of Cnh. One has c n h = Cn x C,.
For even n, the group c n h contains the inversion and can be written as the

Figure A.12: Stereogram of the point


group C B ~ ; ) .
A.3. Tkansfation, point and space groups 649

a) b)
Figure A.13: Stereograms of the point groups C3,(3m) (a) and C4v(4mrn) (b)

direct product C, x Ci of the groups C, and Ci. For odd n, Cnh is equal
toz, as shown in the previous subsection. In any case, the group Cnh is
Abelian. Each element forms a class by itself, thus there are 2n classes. The
stereogram of the group C3h is shown in Figure A.12 as an example.

The group C, may also be complemented by a normal mirror plane uv as


opposed to a horizontal one. Multiplying the powers Ck-', k = 1,2, . . ., n ,
with u,, n different elements U& = Ck-'.uv arise, where awlrepresents u,.
The set of these elements coincides with the set of elements uv.C2-l obtained
by multiplying by C6-l from the right, but the numbering is different. The
elements u,k correspond to n different normal mirror planes. Each adjacent
pair of them forms an angle r / n . For odd n all n mirror planes Uvk are
symmetrically equivalent, for even n they fall into two distinct classes of
symmetrically equivalent planes, the u w l uv3,
, ..,uvn-l with odd k , and the
uvr uv2, .., u,,-~ with even k . One must add all n mirror planes to the
group C, in order to get a new group. The latter then contains, along with
the n rotations C," a total of 2n elements. The international symbol of the
group is nm, if n is odd, and nmm, if n is even. The 'm' following the
odd 'n' describes the system of n symmetrically equivalent normal mirror
planes, and the 'mm' following n describes for the two inequivalent classes of
n/2 such planes. The Schonflies symbol of the group is, uniformly, Cnv. The
group may be understood as the symmetry group of a straight pyramid upon
a regular n-angle. The stereograms of the two point groups C z and C4, are
shown in Figure A.13 as examples. If one replaces the n mirror planes by
n 2-fold rotation axes, then the group C, becomes the gropu D , already
considered. We conclude that the two groups C,, and D , are isomorphic.
650 Appendix A . Group theory for applications in semiconductor physics

a)
i-_)@
0

b)

Figure A. 14: Stereogramsof the point groups D z h ( $ $ $ )


0

(a), and Dsh(6rn2) (b).

The division into classes of D , corresponds to that of C,. This means that
+ + +
C,,, like D,, has 1 (n 1)/2 classes if n is odd, and 3 n/2 classes if n
is even.

Dnh (---,
n 2 2 2nm2)
-
mmm

In this case the group Cnh is complemented by the n 2-fold horizontal axes U2
of D,. Multiplying the 2-fold rotations with Th, one automatically obtains
n normal mirror planes, each of which cuts the horizontal plane f f h of Cnh
along one of the n 2-fold rotation axes. This yields 4n elements altogether.
The group can be interpreted as the symmetry group of a straight prism
upon a regular wangle. The Schonflies symbol reads Dnh. For odd n the
identity = 5 holds. This gives rise to the international symbol z 2 m
in this case. For even n , no uniform relation exists between the subgroup
-n and an rotation-inversion axis. Instead, in this case each mirror plane
m
is normal to one of the 2-fold axes. This leads to the international symbol
$26. Among the subgroups of Dnh are, in each case, the two groups D ,
and C,. Their direct product D , x C, is the full group Dnh. For even n,
D,h contains the inversion. Then the group may also be written as the
direct product Dn x Ci of the two groups D , and Ci. Figure A.14 shows the
stereograms of D2h and D3h.
The division of D,h into classes may easily be obtained from that of D,.
Each class of D , is also a class of Dnh. In addition one has the classes which
emerge from those of D , by multiplying them with f f hor, in the case of even
+ +
n , also with I . This yields n 3 classes if n is odd, and n 6 if n is even.
A.3. Tkanslation, point and space groups 65 1

a) b)
Figure A.15: Stereograms of the point groups D&2m) (a), and D3d(3$) (b).

Consider the division of a prism having the symmetry group D , normally


to the n-fold rotation axis in two half-prisms, and then rotate them towards
each other by an angle xln. In doing so a body emerges whose symmetry
group forms one more point group of the second kind. It includes a 2n-fold
vertical rotation-reflection axis S2,, n 2-fold rotation axes U2 perpendicular
to it, and n mirror planes a d which contain the rotation-reflection axis and
intersect the angle between each pair of adjacent 2-fold axes half way. This
yields, altogether, 4n elements. The Schonflies symbol- of the group is Dnd.
We already know that for even n the relation S2n = 2n holds. The inter-
national symbol is therefore z 2 m . If one adds the inversion to this group,
then the rotation-inversion axis 5 becomes an ordinary rotation axis Cz,,
and the n diagonal mirror planes f f d change over into n diagonal 2-fold axes
U2 by multiplying them with I . A diagonal mirror plane f f d of Dnd, if multi-
plied with the 2-fold axes U2 it contains, yields the horizontal mirror plane
ah. The resulting group is thus nothing but &,h. This means that for even
n we have D d x Ci = D2,h.
For odd n , S2,, = n from which the international symbol fi 6 follows.
For both even and odd n the group D , is a subgroup of Dnd. In the case of
odd n the group contains the inversion (while D n h did so for even n ) , and
may be written as the direct product D , x Ci. The stereograms of the two
particular groups D2d and D3d are shown in Figure A.15.
If one replaces the n mirror planes of D d by n horizontal 2-fold rota-
tion axes Uz in these planes, then the rotation-reflection axis S2, becomes the
rotation axis C2n, and t,he group Dnd becomes the group D2n. These groups
652 Appendix A . Group theory for applications in semiconductorphysics

Figure A. 16: Stereogram of the


point group ~ h ( 6 3 ) .

7
are therefore isomorphic. Knowing this, the classes of Dnd may be obtained
directly from those of D2n. According to what we found earlier, the number
+ +
of classes of D2,, amounts to 2n/2 3 = n 3, whether n is even or odd.
+
The corresponding n 3 classes of D d are E , the n classes {Szn,'"-S,: }, {
sin,s::-'}>,. . . , { SF;', ST$+' 1, { ~ 2 of~powers
) of rotation-reflections
around the two-sided rotation-reflection axis Szn, the class of the n 2-fold
rotations U2, and the class of the n reflections uv.

The group Th may be obtained from the group T of all symmetry rotations
of a tetrahedron by adding the inversion I . The group is therefore the direct
product T x Ci, and contains 24 elements. Besides the inversion these are
the following: In forming the product of the three 2-fold axes C2 with I
one obtains three mirror planes uh perpendicular to these axes. Each of
the four %fold rotation axes gives rise to a 6-fold rotation-reflection axis
sf3 = 3, which means eight new group elements. In Figure A.16 we show
the stereogram of the group. Its international symbol is $3. The classes of
Th follow from those of T : each of the 4 classes of T is also a class of Th, and
in addition there are 4 further classes obtained by multiplying the former by
I . The group T h then has 8 classes altogether.

The full symmetry group T d of a tetrahedron follows from T on replacing


each of the three 2-fold rotation axes C2 by a 4-fold rotation-reflection axis
S4 and adding the six vertical mirror planes uo which contain these axes.
In this way 12 new elements occur, yielding 24 altogether. That the set of
these elements does form a group can be seen from the stereogram in Figure
A.17. The international symbol of the group reads 43m, which reflects its
A.3. Banslation, point and space groups 653

Figure A. 17: Stereogram of the


point group ~ d ( 1 3 m ) .

Figure A. 18: Stereogram of the


point group oh($$$).

construction and is self-explanatory. The group T d is isomorphic to the group


0, where the rotation-reflection axes S4 of T , are assigned to the ordinary
4-fold axes C4 of 0, the uv to the C2, and the C3 again to the C3.
The classes of T d are the 6 elements S4 and Si , the 3 elements ,942 = C:,
the 8 rotations C3 and C z , as well as the 6 reflections uV.Along with E , there
are 5 classes. We abbreviate them by the previously introduced notation
E , 6S4, 3C2, 8C3 and 6uv.

The full symmetry group Oh of a cube may be obtained from the symmetry
group 0 of its rotations by adding the inversion I . Therefore Oh = 0 x Ci.
The 24 new elements include, beside the inversion, the reflections at the
six planes Ud normal to the six 2-fold rotation axes of 0, and the three
mirror planes Uh normal to the three 4-fold rotation axes of 0. The or&-
6 54 Appendix A. Group theory for applications in semiconductor physics

nary rotation axes become rotation-inversion axes, i.e. one has four %fold
rotation-inversion axes 3 in the direction of the %fold ordinary axes of 0,
and three 4-fold rotation-inversion axes in the direction of the 4-fold ordi-
nary rotation axes of 0. The four %fold rotation-inversion axes lead to 8
new elements, and the three 4-fold ones to 6 . The 48 elements of O h are
depicted in the stereograms of the group in Figure A.18. The international
symbol of o h is $3$ (abbreviated as m3m).
The six planes Uh of O h may be interpreted as 6 planes uv which contain
the three 4-fold axes C4. If one observes that each rotation-inversion axis 4
is nothing but a 4fold rotation-reflection axis S 4 , then it becomes evident
how those elements of T d arise in o h , which are not already contained in 0
(these are the four %fold rotation axes and the even-numbered powers S t
of 5'4). From this one may conclude that, with 0, T d is also a subgroup of
o h . Since T d does not contain the inversion, one may understand O h as the
direct product T d x I of T d with I .
The classes of o h may be derived either from those of 0 or those of T d .
The 5 classes E , 6C4,3C;, 8C3 and 6C2 of 0 are complemented by the classes
I , 6 1 . C 4 , 3 I . C:, 8 1 . C3 and 6 I . (32 = Oh. If one refers t o T d , one has first of
all the 5 classes E , 6 S 4 = 6 I . C4,3C,2,8C3 and 6u,, which, by multiplication
with I , then yield the 5 missing classes I , 6 1 . S4 = 6 C 4 , 3 I . C:, 8 1 , C3 and
6 1 . uv = 6C2.
A further point group is the full symmetry group Y h of a pentagondodec-
ahedron. It follows from the icosahedral group Y by adding the inversion.
For the description of the symmetries of crystals, Y h is of as little importance
as Y .
It may be proven that the set of the possible point groups is exhausted
by the above list.

A.3.3 Space groups


Space groups are sets G of rigid displacements g = t a . (Y o f crystals which
transform crystals into themselves. That these sets form groups is proved
as follows. The product g1 . 92 of two displacements g1 = t,, . a 1 and
92 = taz . a2 is again defined by their successive application. Thus, the
equation
91 ' 9 2 = tal+alaz . 02, (A.45)

holds which means that the product is again a rigid displacement. The
crystal will transform into itself under the action of the product g1 . 92,if
it does under the factors g1 and 92 separately. Thus g1 . 92 belongs to the
set G . The other group properties can also be similarly verified using the
product (A.45). The group G represents the spatial symmetry group of a
crystal. If one imagines the crystal as the composite of several superimposed
A.4. Representations of groups 655

lattices of the same Bravais type, then it becomes evident that the lattice
translation group must be a subgroup of the space group.
Theorem 2 of subsection 2.1, wherein each rigid displacement is either a
screw rotation, a glide reflection or a product thereof, is important for the
construction of the space groups. It means that the generating elements of
these groups are screw rotations and glide reflections. Lattice translations
and orthogonal transformations belong to them as special cases. How space
groups may be set up from these elements is explained in the main text.

A.4 Representations of groups

A.4.1 Introduction

The concept of representations

We consider a group G of rigid displacements, say a translation group, point


group or space group. Although representations will be investigated only
for these particular groups, most of our considerations are independent of
particular group properties and hold for groups in general.
A representation of a group is defined in a certain space. Here, we use
the Hilbert space R of quantum mechanical state vectors p of an electron.
Spin will be omitted initially (it will be included later). The components of
a vector 'p are written as functions p(x) of position vector x,where x varies
over the entire (infinite) coordinate space. Since for each x, the value g-lx
is also in coordinate space, p(g-lx) belongs to the Hilbert space if 'p(x)
does. If g runs through the entire group G , then a particular function po(x)
gives rise to as many different functions 'po(g-lx) as there are elements in
the group. Among these functions some number d are linearly independent.
They span a d-dimensional subspace Rd of the Hilbert space R. Let p(x) be
an arbitrary function of R d . Through the relation

9C
pW = (P(g-lx) (A.46)
a h e a r operator j in R d is uniquely assigned to each element g of G . we
will show that the operator 91.92 which according to equation (A.46) cor-
responds to the product g1 . g2 of two group elements g1 and 92,is equal to
the product of the two operators gl and 92, i.e. that the relation

91 ' 92 = Sl . 9 2 (A.47)
is valid. To this end we form ='p(x). By definition we have
656 Appendix A . Group theory for applications in semiconductorphysics

(8.48)
Applying the operator g1 to 92 p(x) = Cp(gF1x), the relation

(A.49)
follows. Comparison of equations (A.48) and (A.49) immediately yields
equation (A.47). This equation means that the linear operators g defined by
equation (A.46) form a group G.
The unique assignment of the original group G to a group G of linear
operators in a certain space Rd, is called a representation of the group G in
Rd. Applying the concept of homomorphy introduced in section 1, one may
also say that a representation of a group G is a homomorphic assignment of
G to a group of h e a r operators in a space Ed. As a general symbol for a
representation we choose D. The space R d is called the representation space.
The dimension d of the space is referred to as the dimension of the represen-
t a t i o n Not every subspace of the Hilbert space R can be a representation
space, although the multiplication rule (A.47) holds everywhere in R. In
order that a subspace Rd have this property, it must contain no vectors p ( x )
such that ~p(9-l~) ( g being any group element) does not also belong to Rd.
The statement that a particular subspace Rd forms a representation space
of a group G is therefore not trivial.
A formal difficulty for the representations of the displacement groups
under consideration, more strictly speaking, for the translation and space
groups, lies in their being infinite groups, i.e. their having infinitely many
different elements. The representations of infinite groups are more difficult
to treat than those of finite groups. Because of this we will focus our consid-
erations on finite translation and space groups. This may be done by means
of the the periodicity region concept introduced in Chapter 2 (the physical
meaning of this concept was also explained there). Instead of the Hilbert
space of arbitrary one-electron quantum states p(x), one uses that of states
being periodic with respect to the periodicity region. Then

(A.50)
holds for arbitrary integers n1, n2, ng. With respect to these states the
translations t R and tnlGlal+nzG2a2+n3G3a3are represented by the same op-
erator ifi. Thus, as long as one deals with representations in the Hilbert
space of functions obeying the periodicity condition (A.50), one can render
the translation and space groups finite by considering only translations t R
through lattice vectors R belonging to the same periodicity region. The
thusly reduced translation group T contains only G3 different elements, as
many as there are different primitive unit cells in the periodicity region.
A.4. Representations of groups 657

The applications of group theory in quantum mechanics rest almost exclu-


sively on the theory of representations. The reason for this is that in quantum
mechanical investigations, as a rule, it is not the groups by themselves which
are of interest, but the operations of their elements on wavefunctions. For
this reason the following theorem, already cited in Chapter 3 is of particular
importance.

Theorem 1

T h e subspace of eigenfunctions of a Hamiltonian having the s a m e energy


eigenvalue f o r m s a representation space of the f u l l spatial s y m m e t r y group
of the Hamiltonian

Below we will demonstrate that so-called irreducible representations are of


particular importance in this context.
Representations depend not only on the groups which are represented
but also on the space which underlies the representation. Here, we have
used the Hilbert space of wavefunctions cp(x) of a single electron without
spin. This gives rise to a particular type of representations (referred to as
vector representations). Later we will consider the space of spin-dependent
wavefunctions cp(x,s ) which introduces representations that do not occur in
normal Hilbert space (so-called spinor representations).
We will now show that the particular representation of the group G of
rigid displacements g defined by equation (A.46) exhibits an important prop-
erty, which other conceivable representations do not have. We are referring
to its unitarity property. An operator g is called unitary if its Hermitian
adjoint operator g+ equals the inverse operator, such that

(A.51)
holds. That these relations are in fact valid for the operators ij defined by
equation (A.46), may be shown as follows. We form the scalar product

(A.52)

of two arbitrary vectors cpl(x) and cp2(x) of Hilbert space R obeying the
periodicity condition (A.50), where the integral is taken over a periodicity
region. The integration variable x in (A.52) is then replaced by g-lx. Since
a rigid displacement does not change the volume element, we find

(A.53)

According to the definition of the hermitian adjoint operator gf one has


(jp2 I g(01) = ( g f . j c p 2 I cpl). Substituting this relation in (A.53), it follows
658 Appendix A . Group theory for applications in semiconductorphysics

(cp2 I cpl) = (3+. 3 9 2 I cpl). (A.54)


Because 9 1 and cp2 are two arbitrary vectors of Hilbert space R , the relation
(A.51) must hold in R.

Representation matrices

It is convenient to describe the operators g of a representation as matrices


with respect to a particular orthonormalized basis set cp', p 2 , .. . cpd in the
representation space R d . Since the vectors, $@
+ gcp2,.. . gpd are likewise vec-
tors in E d , they may be written as linear combinations of basis functions
cpl, p 2 , .. . cpd, whence

(A.55)

with Dpl(g) as expansion coefficients. The Diq(g) are elements of a quadratic


matrix D(g) with d rows and d columns. The operators 9 are thus described
by d-dimensional square matrices D(g). Since the 3 form a representation of
the group G, the same also holds for the matrices D(g) representing them.
The 'product operation' between the elements of the matrix group is ordinary
matrix multiplication. This may be seen by calculating the matrix D ( m )
of a product element 91.92.By definition, we have

(91 .92) cpZ(X)= cDZIl(g1 . g2)cpZI(x). (A.56)


I'
On the other hand. it is also true that

(A.57)

Comparison of equations (A.56) and (A.57) yields

(A.58)

D(g1 ' 92) = D(g1) ' D(92). (A.59)


The properties of the representation operators g are carried over to the
representation matrices D ( g ) . Thus, the unitarity of the operators 3 results
in unitarity of the matrices D(g) such that
A.4. Representations of groups 659

D+(g) . D(g) = 1 or D+(g) = o-'(g). (A.60)

The Hermitian adjoint matrix D'(g) is defined as the complex conjugate


transpose matrix [oT(g)]*.

Special representations. Equivalence

Each group allows for several more or less obvious representations. One of
them is the so-called identity representation.

Identity representation

Every group element is associated with the identity operator or the identity
matrix, in this representation space. The 1-dimensional identity representa-
tion can be constructed as follows: One considers an arbitrary vector p(x) of
Hilbert space, and forms the new vector cgl p(g'-'x), where the summation
runs over all elements of the group G . This vector is then a 1-dimensional
representation space of G, and the representation in it is the identity repre
sentation. To prove this statement we apply an arbitrary element g of the
group to Cgl p(g'-'x). In this, the sum reproduces itself because g'-lg-l
runs through all group elements if g varies over the entire group. Thus we
may write

(A.61)

According to this relation, all g are identity operators in the 1-dimensional


space cgl p(g'-lx), and the representation is indeed the 1-dimensional iden-
tity representation, as claimed. It is also now clear how one can construct
identity representations of higher dimensions. One selects linearly indepen-
dent vectors cpl(x),cp2(x) etc. in such a way that the sums Cgl cpI(g'-'x),
Cgl(02(g'-lx) etc. are linearly independent of each other. Then the vectors
Cglpl(g'-'x), Cgl'p2(g'-'x) etc. span a space in which all vectors trans-
forms into themselves under the action of all elements of G , in other words
it forms a multi-dimensional identity representation.
From a known representation V ,additional representations of the same
dimension can be derived. Below, we give the most important ones.

Complex conjugate representation V *

This representation is obtained when the representation space Rd of V is


replaced by the complex conjugate space Rfi. The operators g* representing
an element g in the representation V*are defined by the equation
660 Appendix A . Group theory for applications in semiconductor physics

g*cp*(x)= q"(g-1x). (A.62)


The matrices D*(g)of the complex conjugate representation D* are the
of the original matrices D(g),
complex conjugate matrices [D(g)]* whence

= [D(g)l*.
D*(!?) (A.63)
A representations is called real if its matrices are real, i.e. if

(A.64)

-
Conjugate representation D

The conjugate representation 5 of a representation D is defined by assigning


matrices b(g)to the elements g given by the equation

= [D(g-l)lT. (A.65)
One can easily demonstrate that the matrices b ( g ) form actually a repre-
sentation if the matrices D(g)do so. In the case of a unitary representation,
the conjugate representation equals the complex conjugate representation
because

(A.66)

Equivalent representations

Transforming the representation space Rd by means of an arbitrary non-


singular operator M likewise results in new representations of the same
dimension. A representation D M is said to be equivalent to another rep-
resentation V if it is associated with the group G M of operators

jM - M . j . M-1. (A.67)
The group property of G M follows directly from that of the original operator
group G. The associated matrices DM(g)of the equivalent representation
read

DM(g)= M . D(g). M - l , (A.68)


where M denotes the representation matrix of the transformation opera-
tor A?. Since A? can be any non-singular operator in the representation
A.4. Representations of groups 661

space R d , one can produce infinitely many equivalent representations from


a given representation 2). It turns out that it is often unimportant to make
a distinction between equivalent representations, and that only inequivalent
representations are to be considered as essentially different. Therefore, one
often refers to equivalent representations as representations which are equal
apart of equivalence. Occasionally, one omits the qualification apart from
equivalence and just calls equivalent represent ations equal.

A.4.2 Irreducible representations

Reducibility and irreducibility

In the above definition of a group representation we did not exclude the pos-
sibility that the representation space R d contain several genuine subspaces
which are also representation spaces. If this is the case, the representation
space is called reducible and the representation of the group in it is called a
reducible representation. If the representation space does not contain such
genuine subspaces, which are also representation spaces, one calls it irre-
ducible. The representation of a group in an irreducible space is called an
irreducible representation. The following important theorem can be proved:

Theorem 2

Each reducible unitary representation can be decomposed i n t o irreducible uni-


tary representations.

The proof of this theorem will be briefly sketched below. By definition, in


every reducible representation space Rd there is at least one subspace which
is itself a representation space. Let R1 be such a subspace, either reducible
or irreducible. Using the unitarity of the representation one can easily show
that the complementary subspace %of R1 in R d is orthogonal to R1. Under
the operations of a unitary representation the complementary subspace
is then transformed into a subspace of Rd which is again orthogonal to R1.
Thus is mapped into itself, and it forms a representation space contained
in Rd as well. If a smaller representation space R2 also exists in K, then its
complement in K also forms a representation space contained in R d . One
can continue this procedure until one reaches a remainder space which no
longer contains a genuine subspace which is also a representation space. This
space is then, by definition, irreducible. The same procedure can be applied
to the representation spaces R1, R2 etc., which one has chosen in decompos-
ing R d into subspaces. If these paces are reducible at the outset, then one
can also decompose them into irreducible representation spaces by means of
the above procedure. In this way a complete decomposition of the originally
662 Appendix A. Group theory for applications in semiconductor physics

considered reducible representation space Rd into irreducible representation


spaces R,,, v = 1,2,. . . has been accomplished. The corresponding repre-
sentations D,,are irreducible. The original reducible representation D is
decomposed, therefore, into a number of irreducible representations D,,and
the theorem is proved.
The decomposition may be visualized by means of the representation
matrices D ( g ) of a reducible representation D. We consider a basis in the
representation space R d , which is composed of basis sets of the irreducible
representation subspaces of R d . In this basis the d-dimensional matrices
D ( g ) of the reducible representation Ig decompose into the matrices D,(g),
D,t(g), D,tt(g), . . ., of the irreducible representations D,, D,,, D,tt, . . . con-
tained in D. The matrices D,(g), D y , ( g ) ,D,,t(g), . . . are situated along the
diagonal of D ( g ) as shown below:

\ ..... 1
One says that the matrices D ( 9 ) are block-diagonalized. The block-diagonal
form is thereby the same for all g , i.e. the matrices D,(g), Dvt(g), D,,i(g), . . .,
for all group elements are, respectively, situated at the same positions on the
main diagonal, and have, respectively, the same dimensions.
An irreducible representation V, exists for a subspace R, of Hilbert
space such that, firstly, every g maps R , into R, (which means that R,
indeed forms a representation space) and secondly that there is no yet smaller
nontrivial subspace of R , with this property. We already know a special 1-
dimensional irreducible representation, namely the 1-dimensional identity
representation. From this we conclude that each group has at least one
irreducible representation. In general there may be several.
That the degenerate eigenfunctions of a particular one-electron Hamilto-
nian having the same energy eigenvalue form a representation space of the
full symmetry group of the Hamiltonian, was already stated in theorem 1.
If the energy eigenvalues are degenerate only because of spatial symmetry,
i.e. if no accidental degeneracy exists due to the particular values of the
potential or to time reversal symmetry, then it is possible to make an even
stronger statement:

Theorem 1'
The subspace of eigenfunctiom of a given Hamiltonian having the same en-
ergy eigenvalue is a representation space for an irreducible representation of
the f i l l spatial symmetry group of the Hamiltonian
A.4. Representations of groups 663

We will now discuss important relations which hold for matrices of irreducible
representations.

Orthogonality relations

The elements D $ ( g ) of the representation matrices may be understood as


complex scalar functions defined on the group. In analogy to the ordinary
scalar product between vectors we define the scalar product of two such
arbitrary functions y ( g ) and 6 ( g ) through the relation

(A.70)

The factor 1/N with N being the number of elements of G is added for
reasons of normalization. Orthogonality in the sense of this scalar product
has the usual meaning (6 1 y ) = 0. With this definition, we have the following
theorem:

Theorem 3

The matrix elements Drk(g) and Dt",(g) of two inequivalent unitary irre-
ducible representations are orthogonal functions over the group, i.e. they
obey the relation

(A.71)

If the two representations Dp and V" are equivalent, then the matrix ele-
ments D & ( g ) and D z k , ( g ) are orthonormal functions, such that

1
(DCk, I DU) - - 6ij~6kkf,
-d,
(A.72)

with d, = d, as the common dimension of the two representations.


We know that the 1-dimensional identity representation D1,which as-
signs each group element to the number 1, forms an irreducible representa-
tion for any group. If one identifies the representation D, with V1 in relation
(A.72), then the equation

(A.73)

follows.
Important instruments for investigation of the properties of representa-
tions are their characters. These will be treated now.
664 Appendix A. Group theory for applications in semiconductor physics

Characters of representations

A d,-dimensional irreducible representation D, is composed of N matrices


D'(g) of dimension d,. The sum of diagonal elements of a matrix D , or
'trace'

(A.74)

has the property that it does not change under a unitary transformation of
the basis with a given matrix U . This is clear if we recall that the matrix
D in the new basis is given by U-IDU and that under the trace symbol
'Tr' the factors of a product of two matrices can be interchanged. We have,
therefore, TrIUDU-l] = T r ( U - l U D ] = T r [ D ] . This property gives rise to
the definition of the character of a representation in the following way. By
means of the relation

(A.75)
each element g is assigned a particular complex number X,(g). This number
is called the character of the element in the representation D,. The totality
of the character values X,(g) for the various elements g of a group constitute
the character X , of the representation One can thus say that the character
is a complex scalar function defined on the group. The character of the d-
dimensional identity representation has the constant value d over the whole
group.
Two different elements g' and g", which are conjugate to each other in
the sense of section A. 1, i.e. such that there is an element g of the group for
which gN = gg'g-' holds, have the same character X,(g') = X,(g"). This
follows, again, because of the commutativity of the operators under the trace
symbol. One can also say that the character function X,(g) has a uniform
value for each class of conjugate elements. The values change only in passing
between different classes.
In subsection A.4.1, starting from a given representation, a series of other
representations was constructed, among them the complex conjugate repre-
sentation, the conjugate representation and the equivalent ones. The char-
acters of these representations bear well defined relationships to those of the
original representation. For the character X v * ( 9 ) of the complex conjugate
representation D: of a given representation D, one obtains

(A.76)
Real representations have real characters. From the reality of the character
one cannot, however, deduce the reality of the representation, since complex
representations can also have real characters.
A.4. Representations of groups 665

For the character X c ( g ) of the conjugate representation 6uof a given rep-


resentation Vurit follows that

(A.77)

For unitary representations this relation can also be written as

Thus the character X , ( g ) of an element g must be real in such a represen-


tation if g is in the same class as the inverse element g-'.
The characters X,M (9)of all equivalent representations D,M of a given
representation D, are equal and coincides with the character of V,. Thus,

If the complex conjugate representation V * is equivalent to the representa-


tion V itself then the character of D must be real.
Relation (A.79) is, again, a consequence of the commutativity of matri-
ces under the trace symbol. That equivalent representations do not differ
by their characters underlines the above-made statement that there are no
essential differences between such representations.

Theorems on irreducible representations

There are a number of important theorems relating to the irreducible rep-


resentations of a group. Here, we will discuss the ones which are of direct
relevance for semiconductor physics and which are used in the main part of
the book.

Theorem 4

A finite group has exactly as m a n y inequivalent irreducible representations


as it has classes of conjugate elements.

It follows that finite groups can have only a finite number of inequivalent
irreducible representations. In all cases, one of them is the 1-dimensional
identity representation.

Theorem 5

T h e s u m of the squares d: of the dimensions of all non-equivalent irreducible


representations D v yields exactly the number N of the elements of the group,
so that
666 Appendix A. Group theory for applications in semiconductorphysics

Ed: = N . (A.80)
U

This statement is referred to as Burnside's theorem. An immediate conse-


quence is that the dimension d, of any irreducible representation of a finite
group must also be finite. There is an even stronger statement relating to
dv :

Theorem 6

The dimension d , of any of the irreducible representations of a finite group


i s a divisor of the group order N .

Using theorems 4, 5 and 6, and considering the fact that among the irre-
ducible representations there is always a 1-dimensional representation (namely
the identity representation), one often can determine the dimensions of the
irreducible representations of a group without knowing the representation
matrices themselves. The group &d, for example, has 8 elements and 5
classes. There is only one way to obtain the group order 8 by adding 5
+ + +
squares, namely by taking four squares of 1 (12 l2 l2 l2= 4),and one
of 2 (22 = 4). This means that the group D w has four 1-dimensional and
one 2-dimensional irreducible representations. For an Abelian group Burn-
side's theorem immediately implies that only 1-dimensional representations
are allowed for in such a groups each element forms its own class, so that
the number of classes equals the number N of group elements. Because of
(A.80) then all d , must then be 1.

Theorem 7

The characters of the irreducible representations of a group are orthonormal


with respect to each other, whence

Since one of the irreducible representations is the 1-dimensional identity


representation D1,whose character has the value 1 for all group elements,
we obtain from equation (A.81)

(A.82)

By means of this theorem the characters of the irreducible representations


A.4. Representations of groups 667

can be often already determined without knowing the representation matri-


ces.
Each unitary reducible representation V can be decomposed into irre-
ducible representations, as was shown at the outset. A particular irreducible
representation may occur repeatedly. If m, is the multiplicity of the irre-
ducible representation V, in the reducible representation V ,then the char-
acter X(g)of D may be written in the form

(A.83)

where the sum is extended over all distinct irreducible representations V, of


the group. Using the orthonormality relation (A.81), one can determine the
multiplicities m,r immediately. Upon scalar multiplication of X with Xu(,
it follows that

(xV/I X )= x(g)= Cm,(x,/ I x,) = m y / . (A.84)


Y

This relation provides a simple procedure for determining if and how often
a particular irreducible representation V, occurs in a supposedly reducible
representation 2). The reduction of V into VU1smay be written in the
symbolic form

= Cm,D,. (A.85)
V

The summation in (A.85) is to be understood in terms of placing mu copies


of the dv-dimensional matrices on the main diagonal of the d-dimensional
matrix D. One has, of course, d = Cumu . d,.

A.4.3 Products of representations

Definitions and theorems

We now introduce the concept of the direct produet of two representations.


Let R, be a representation space for the representation V, of G, and Rv
be a representation space for V, of G. We will show that the product
space R , x Ru of the two spaces is also a representation space. Let cp,(x)
be an arbitrary vector of R,, and cp(x) an arbitrary vector of R,. Then
cpf(x) . cpV(x)forms a general vector of the product space. Since, with
v p ( x ) .cp(x) we also have yP(g-lx) . q ~ ( g - ~ xin )the product space, the
equation

g cp,(x). cp(x)= yP(g-1x) . (p(g-1x) (A.86)


668 Appendix A. Group theory for applications in semiconductor physics

assigns each element g of the group an operator g. This assignment is a


representation of the group in the product space, called the product repre-
sentation D, x V,. Along with (A.86) we also have

g p ( x ) . p,(x) = p ( g - 1 x ) p y g - l x ) (A.87)
The direct product representation is therefore independent of the ordering
of the factors, i.e. we have

D,xV,=V,x~,. (A.88)
The matrices D P X u ( g )of the direct product representation V, x D, can
be easily calculated from the matrices D p ( g ) and D ( g ) of the two fac-
tors. To this end we introduce a basis in the product space. Let the vectors
p y ( x ) ,p g ( x ) ,. . ., p$@(x)
be a basis in R,, and the vectors pi(x),p ? ( x ) ,. . .,
p z p ( x )are a basis in R,. As a basis of R , x R,,we therefore have the prod-
ucts pf(x) . p;E(x) with 1 = 1,2,. . . ,d , and k = 1 , 2 , . . . ,d,. By definition,

D:tlyk = D:l . DLfk. (A.91)


For the character X P x , of the direct product representation it follows from
(A.91) that

or

X,xv(g) = &(g) .X d g ) (A.93)


This is summarized in the following theorem:

Theorem 8

The character of the direct product of two representations is the product of


the characters of the factors.
A.4. Representations of groups 669

The direct product representation is in general reducible even if the factors


were irreducible. To find all irreducible parts one may use relation (A.84).
It is of particular interest to know whether the identity representation is
contained in the direct product or not. The answer to this question is given
by the following theorem:

Theorem 9

T h e direct product Vz x V, of two unitary representations V*, and V, con-


tains the identity representation exactly i f and only if the representations V p
and V, are equivalent t o each other. T h e identity representation then occurs
exactly once.

We will now consider the direct product of a representation with itself.

Direct product of a representation with itself

In this case one has

The product space R u x , can always be decomposed in this case into two
representation subspaces R t x u and R;,,, where RExu is spanned by the
symmetrical combinations (l/a)[Cp;(x) . $(XI) +
$(x) . (p;(x)] of the
products of basis functions, and RExu by the antisymmetric combinations
(l/&)($(x). cpi(x)- (pi(x) . &(x)]. The dimension of REx, is given by
d(d f 1)/2, and that of RExu by d(d - 1)/2. The representation [V,x V,I8
of the group G in RExu is called symmetric product and the representation
[V,x V,],in RExu is the antisymmetric product. For the matrices DL:Jc(g)
of [Vvx V,],we find from (A.89) and (A.90),

(A.95)

and for those of [V,x V,],,


we have

(A.96)

The pertinent characters X i x u and XExu are

(A.97)

(A.98)
670 Appendix A. Group theory for applications in semiconductor physics

Selection rules for matrix elements

We consider a particular group G of rigid displacements. In practical cases,


G is the symmetry group of the Hamiltonian. The following considera-
tions are not limited, however, to this meaning of G. Given two irreducible
representations D ' , and D' , of G , we take cp';(x),( O ~ ( X .) ., . ,cpzv(x) as a set
of basis functions in the representation space R , belonging to V,, and
cpy(x),& ( x ) , . . . ,cpZp(x)as a basis set in the representation space R , be-
longing to V,. In the case considered above, the v y ( ~&'(x), ) , . . . ,cp$Jx)
and cpy(x),&(x), . . . , & J x ) form complete sets of eigenfunctions of the
Hamiltonian for particular energy eigenvalues E , and E,, respectively. We
will examine the matrix elements of a linear operator A in Hilbert space
with respect to these basis functions. The operator A may be a tensor of
arbitrary rank. We denote its components with respect to a particular basis
e l , e 2 , .. . of the tensor space by A1, A 2 , . . .. The basis vectors e k are tensors
of the same rank as A. To be able to use the theory of representations, we
must assume that the components A k transform according to a particular
representation DA of G. Thus, we should have

(GAg-')k = D$kAkl. (A.99)


kf

For a scalar operator A , relation (A.99) takes the form

A. (A.100)

If G is the symmetry group of H , then H itself can be taken as scalar


operator. For the components ( A 1 , 6 2 , A g ) of a vector operator A with
respect to a Cartesian coordinate system e l , e2, e3, one obtains from (A.99)
the transformation

(A.lO1)
k'

with Dgfk(g)as matrices of the representation VV of G which governs the


transformation of 3-dimensional vectors. The momentum operator p serves
as an example of a vector operator.
w e now form the matrix elements of the operator components ( c p t I A k 1
(py) of A between the basis functions (pr(x),1 = 1,2, . . . , d,, and &(x), m =
1,2,.. . , d,. Because of the unitarity of the operators g, we have

(A.102)

or, considering equation (A.99),


A.4. Representations of groups 671

This means that the matrix elements (&&l(Pr) transform according to the
direct product representation 23; x V Ax D,. Summing over all elements g ,
we obtain from (A.103) the relation

(A.104)
If the identity representation does not occur among the irreducible parts
of the direct product representation V; x DA x D,, then the sum over g
vanishes according to relation (A.73). In this case, one correspondingly has

i.e. all matrix elements ($7kl&(pfl) vanish for symmetry reasons. If the
identity representation is contained in the direct product representation 21: x
DA x D p , the sum over g in (A.104) does not vanish and the matrix elements
($&1&1$7r) can be non-zero. Nevertheless, they may vanish for reasons
unrelated to spatial symmetry.
If some of the elements ( & [ A k l p r ) do not vanish by reason of symmetry,
questions remain open as to which of the elements are non-zero and what
symmetry-induced relations exist between them, in other words, how many
independent non-zero matrix element exist. We now proceed to discuss this
question. To this end we start with equation (A.103) and rewrite it using
the representation matrices

(A.106)

pxAxu
($7hI A k I $7); = Dmklmkl(g)($7& I AkIpF). (A.107)
lmk

Because of the reducibility of the direct product representation D;x V Ax D,,


there is a unitary transformation U of the basis [cpf(x)]* . e k . cpL(x) of the
product space such that the matrix D$j$Akl(g) in the transformed basis
takes the block-diagonalized form (A.69). We denote the transformed basis
by [&(x)]*. e K . cp&(x),and the matrix elements with respect to this basis
~XAXU
by D M ! K i L i M K L ( g ) . Then We have
672 Appendix A. Group theory for applications in semiconductorphysics

where the DLtKtL!MKL(g) denote the block-diagonal matrices in the product


space. All their blocks are zero except for one. The one non-vanishing
block is the representation matrix d i ) ( g ) of an irreducible representation
V A ( i ) of G . For different i, X ( i ) can have the same value, i.e. a particular
irreducible representation matrix D x ( g ) can appear as a block in several
matrices D L I K , L I M K L ( g )The
. number of these matrices is the multiplicity
mA of equation (A.84) with which Vx is contained in VoT/x V Ax V p .
To facilitate the use of the decomposition (A.108) in relation (A.107), one
must transform relation (A.107) into the new basis. In this, the summation
over all elements g of the group G yields

(A.109)
Because of the orthogonality relation (A.73), the g - s u m results in a non-zero
value only for i such that Vx(i)is the identity representation. We assume
that this occurs in the matrices DLlKtLiMKL ( 9 ) for i = 1 , 2 , .. . ml. For
these values of i, the DL,,,,,,,,(g) are diagonal matrices with only one
non-vanishing element. The one non-zero element has the value 1. We denote
the position M'K'L' = M K L of this element by MiKiLi. Then equation
(A.109) may be written as

Since 1 5 i 5 m l , these relations mean that exactly ml matrix elements


( c p L l A ~ I p ; )can be different from zero. The matrices (&IAklpy) with re-
spect to the original basis follow from (pLlA~lp;) by means of the transfor-
mation U + inverse to U . In the back-transformed matrix, which is the orig-
inal matrix (pklAklpr), generally, all elements (rpklAk1pf)differ from zero,
but they all depend linearly on the mi non-vanishing elements ( p & M ( A ~ ( p z )
of the transformed matrix. Altogether, there are m i independent non-
vanishing matrix elements (&I A k l p r ) . By means of the unitary transforma-
tion U ,the matrix (pk,lAkJ(~f) can be calculated explicitly from the known
matrix ( p L l 2 ~ l p L ) .The results above are summarized in the following
theorem:
A.5. Representations of the full rotation group 673

Theorem 10

T h e matrix elements (p$lAklp[) of the operator vanish f o r s y m m e t r y


reasons i f and only if the direct product representation VE x V A x V, does
n o t contain the identity representation. If the identity representation i s in
VE x V Ax V,, n o t all elements (p&]Aklp[) vanish by reason of s y m m e t r y .
There are as m a n y independent elements among t h e m as the number of t i m e s
the identity representation i s contained in Vi x V A x D,.

From theorem 9, and the commutivity of the factors of a direct prod-


uct representation, we conclude that the identity representation appears in
VE x V A x V v as often as the representation product V; x V, contains the
irreducible representation VA.

A.5 Representations of the full rotation group


Although the full rotation group has no direct meaning for crystals, indi-
rectly, however, it plays an important role in their study. Its representations
define the space of electron states in which the operators of quantum me-
chanics act and, consequently, in which the representations of the symmetry
groups of crystals need to be considered.
The full rotation group is an infinite continuous group. This distin-
guishes it from the groups considered heretofore, which were finite and dis-
crete. Many concepts and theorems formulated for the latter can be utilized
directly for infinite groups. Among them is the group definition itself and
the concepts of conjugate elements and classes of such elements. In the
full rotation group, all rotations through the same angle, but with respect
to different axes, are conjugate to each other. The corresponding classes,
therefore, contain an infinite number of elements. The number of different
classes is also infinite because the number of the different axes is infinite.
The concept of representations may be used for the infinite rotation group
as well. The dimension of the representation matrices is, however, no longer
necessarily finite, as in case of finite groups. Matrices of any size are allowed,
among them, of course, also finite ones. The reducibility and irreducibility
of representations are defined in analogy with the corresponding definitions
for finite groups.
A number of theorems which were formulated for finite groups lose their
validity, however, or must be modified when applied to infinite groups. That
this happens is then evident, if the group order or the number of classes
enters the theorem directly. For example, one can only speculate whether
the number of inequivalent irreducible representations of the rotation group
is actually infinite, as it follows formally from theorem 4 and the fact that
the number of classes is finite, for it is not clear whether this theorem still
6 74 Appendix A. Group theory for applications in semiconductorphysics

holds in the case of infinite groups at all. We shall see below that there
is in fact an infinite number of inequivalent irreducible representations of
the rotation group, so that the formal application of theorem 4 is justified.
Generally, one ought to use care in transferring results for finite groups to
infinite ones. Luckily, it turns out that the construction of the irreducible
representations of the rotation group (being the aim of this section) does not
require the theorems derived above.
O n e of the irreducible representations of the full rotation group we know
already, namely that in the 3-dimensional vector space, which is called the
vector representation The corresponding representation matrices follow
from the matrix A of equation (A.31), if the Eulerian angles are allowed
to take all real values. The matrix A represents rotations in a special ba-
sis, namely that of the Cartesian coordinate system before the rotation. As
we will see, there exists in each representation space, also including that of
dimension 3, exactly one (inequivalent) irreducible representation. Each 3-
dimensional irreducible representation may, therefore, be obtained from the
matrix A of equation (A.31) by a unitary transformation. The reason that
we will once again deal with the 3-dimensional vector representation is that
this representation is suitable for introducing a concept which is of impor-
tance for irreducible representations of any dimension, namely the concept
of generators of infinitesimal rotations.

A.5.1 Vector representation of the rotation group and gen-


erators of infinitesimal rotations
The following construction of the vector representation of the rotation group
G R relies on the fact that each rotation can be described by a directed axis
and a rotation angle cp with respect to this axis. We denote the unit vector
in the axis-direction by n. Let the Cartesian unit vectors be e,, ey,e,.
We maintain that the coordinates x,y, z of the rotated vector x may be
obtained from the coordinates 2,y, z of the vector x before rotation by means
of the relation

(A.111)

(A. 112)

is the rotation matrix. The components Iz,Iyl I, of I are called generators


of the infinitesimal rotation. They are given by
A.5. Representations of the full rotation group 675

0 0 i 0 -i 0
I, = 0 0 -i , Ip= 0 0 0

O i 0 -i 0 0 0 0 0
113)
The exponential function in (A.112) is to be understood in the sense of a
power series expansion with respect to -i(I. n)p. In the following we sketch
the proof of assertions (A.lll) and (A.112).
Initially, we consider a rotation about the z-axis through the infinitesimal
angle dp. The coordinates after rotation, obtained using the matrix A of
+
(A.31), are given by 2 = 2,y = y - zdq,, z = z ydp,. In matrix form
these equations read

(A.114)

with
D(ex,dp,) = E - iIxdvz, (A. 115)
and with E the 3-dimensional identity matrix. Analogous relations hold for
infinitesimal rotations with respect to the y- and z-axes. If the three in-
finitesimal rotations are executed one after another, then this corresponds,
apart from higher order corrections, to a rotation about an axis in the direc-
tion of the vector d$ = e,dp,+eydpy+e,dp, . With n as unit vector in the
+
direction of d$, d p as the absolute value of d$, and I = Izex Igey Izez, +
the result of the rotation can be written as

(A. 116)

with
D(n, d p ) = E - i ( I . n)dp. (A. 117)
Now consider a rotation through the finite angle cp with respect to the same
rotation axis n. The pertinent rotation matrix is D(n,p). After this ro-
tation we apply the infinitesimal rotation (A.117). The total rotation is
+
described by the matrix D(n, p dp). Because of the group property of the
representation matrices, we have
676 Appendix A . Group theory for applications in semiconductor physics

On the other hand, expanding D(n, cp+dcp) with respect to the infinitesimal
angle dcp, we obtain

(A.119)

Comparing with (A.117) it follows that

d
-D(n, cp) = - i ( I . n)D(n, p), (A. 120)
dcp
This differential equation is to be solved with the initial condition D(n, 0) =
E. The result is just expression (A.112) for D(n, cp).

A.5.2 Representations for dimensions other than three


Equation (A.112) relates the rotation matrix with the three generator ma-
trices I z , Iy,Iz. Through the latter D(n, cp) is uniquely determined. If, in
turn, the rotation matrix D(n, cp) is known, the three generator matrices can
be determined from it as follows

d d d
I , = i--D(ez, cp)lrP=0, Iy= i-D(eY, cp)lrPP=o,I , = i-D(e,, cp)lrP=o.
dcp dP dcp
(A.121)
These relations may be understood as the defining equations of the gen-
erators matrices. As such, they may also be applied to representations of
dimensions other than 3. With this in view we shall consider D(n, cp), and
I%,Iy,Iz as matrices of any dimension henceforth. With a knowledge of the
generator matrices for dimension d, one can then determine the rotation
matrices of a d-dimensional irreducible representation by means of equa-
tion (A. 112). In this way, the construction of the irreducible representations
of the rotation group is traced back to the determination of the generator
matrices. To accomplish the latter task, the commutation relations among
the matrices are of great value. In the 3-dimensional case the explicit form
(A.113) of the matrices I,, Iy, I, may be employed to show that

[I,, Iy] = iI, , Iz] = iI, , [Iz,I,] = iIv.


[Iyr (A.122)
In quantum mechanics, the same commutation relations hold for the com-
ponents of the angular momentum operator. This reflects a specific relation
between the two sets of quantities, which we will discuss further below. The
definition (A.121) can be used to demonstrate that the commutation rela-
tions (A.122) apply to generator matrices of any dimension. One may thus
use them for the construction of these matrices. Before we can do this, we
must address two questions which, thus far, have been ignored the questions
A.5. Representations of the full rotation group 677

which dimensions are allowed for the irreducible representations, and how
many inequivalent representations exist for a given dimension.
Consider an irreducible representation 2) of the rotation group GR in a
d-dimensional space Ed. Being a representation space of GR, Rd will also
simultaneously be a representation space of the subgroup G , of all rotations
about the z-axis. Due to the Abelian type of G,, its representations are
1-dimensional. Thus, D considered as a representation of G,, falls into a
series of 1-dimensional representations. Let the corresponding basis vectors
be b l , b2, . . . , bd. we will relate the rotation matrices D of the representation
D to them in the following. For a rotation through an angle cp with respect
to the z-axis. we have

D ( e , , cp) b, = e-amuv b,, Y = 1,2,. . . ,d, (A. 123)


with m , an arbitrary integer (including zero). Differentiating relation (A. 123)
with respect to cp and applying equation (A.121), it follows that

I,b, = m,b, , Y = 1,2,. . . , d. (A.124)


Instead of the generator matrices I , and I y , it is expedient to use the two
linear combinations

I+ = I , + iIy, I- = I, - iIy. (A.125)


By j we denote the largest of the integers m l , m2,. . . ,mu. Using of the
commutation relations (A.122), it may be shown that the vectors

ej-1 = I-bj , ej-2 = I-ej-1 , ej-3 = I-ej-2 , ... (A.126)

are all eigenvectors of I,. They satisfy the equations

Ire, = m e , , m =j,j - 1,j- 2 , . .., (A.127)


where for the sake of uniformity we have set e j bj. The ej-k are mutually
orthogonal. Also they all lie in the representation space R d , so that not more
than d of them can be different from zero. For a particular value k = T , we
must therefore have

ej-T = I-.bj-T+l = 0. (A.128)


Similarly, by using the commutation relations, one can show that the oper-
ator I' = I," i-I: +I," satisfies the eigenvalue equation

(A.129)
678 Appendix A . Group theory for applications in semiconductor physics

with E as the identity matrix in Rd. To prove this relation, one first verifies
the equation
I+ej = 0 (A.130)
arguing indirectly: If the vector I+ej were to be non-zero, then it would
be an eigenvector of I , having eigenvalue j +1. We have already assumed,
however, that j is the largest eigenvalue of Iz in R d . On the basis of this
contradiction, the assumption that relation (A. 130) does not hold must be
rejected.
Using this result, one first proves equation (A.129) for m = j , by express-
ing Ix and I, in terms of I+ and I-. To do the proof for arbitrary values
+ +
of m , equation (A.126) and the commutativity of I," I; I,"with I- are
sufficient.
Equation (A.129) makes it possible to calculate the effect of I+ on the
basis vectors em with m # j . To this end, one replaces em in terms of
I-e,+l. +
For I+I-, the relation I+I- = I 2 I, - I,"may be employed, with
the result

I+em = [ j ( j + l ) - m(m +l)]em+l , m =j - 1 , j- 2 , . . . , j - r + l . (A.131)


Owing to above derived relations it is now clear that the matrices I+, I-, Iz
transform the space spanned by e j , e j - 1 , . . . , ej-,.+l into itself. The same
holds, of course, for the generator matrices Ix,Iy,I, and the matrices D
of the irreducible representation 2, of the rotation group defined by them.
According to the initial assumption, the dimension of this representation is
d , whence we have r = d.
There is an intimate relation between d and j , which may be seen as
+ +
follows. The diagonal matrix elements (em I 1," I; I," 1 em), may be
evaluated using equation (A.129), with the result

(em I 12 I em) = j ( j + 1) = m2 + (em I I: +I; I ern). (A.132)

Since here (em I I: +


1; 1 em) is positive, then m2 < j ( j + 1). The largest
negative integer which satisfies this condition is m = - j . Thus m takes the
+
(2j 1) values - j , - j +
1,. . . ,j - 1, j , and consequently d = 2 j 1, or,+
j = -
d-1
(A. 133)
2 -
Thus j is a half integer if d is even, and an integer if d is odd.
Relations (A.126) to (A.131) define the elements of the matrices I+, I-, 1,
and, therefore, also those of the generator matrices Iz,Iy,I, with respect
to the basis e j , ej-1, ..., e - j . The disadvantage of this basis is that it is
not yet normalized, so that the generator matrices, and the corresponding
A.5. Representations of the full rotation group 679

representation matrices are not yet unitary. Therefore, we now normalize


the em and denote the normalized basis vectors by 2., The latter obey the
following relations in place of (A. 126):

im = N m I-6,+1 (A.134)
with
1
Nm = (A.135)
J(j + m + l)(j- m )
as normalization factor. The elements of the matrices I+,I - and I, in the
normalized basis become

(Em I Iz 1 i m l ) = m6,1. (A.137)


The generator matrices of a given dimension result in an irreducible repre-
sentation of the rotation group of the same dimension. This is evident from
equation (A.112), which determines this representation uniquely. One can
show that this is also the only irreducible representation of this dimension.
An irreducible representation of the full rotation group is thus uniquely de-
termined by its dimension d or, in view of equation (A.133), by the parameter
j. One usually employs j and denotes the various irreducible representations
by Dj. The corresponding representation matrices are denoted by D j . Their
elements (im I D j 1 Cm,) may be calculated by using equation (A.112), as has
already been discussed, but other, methods based directly on the differential
equation (A.120) are more practical. We will not discuss this matter here in
detail, but only give the final results. Using Eulerian angles for the rotations
a,one finds

3 - m ) ! ( j+ m')!
X

with p = cos 0. The following relations may be derived:


680 Appendix A. Group theory for applications in semiconductor physics

(im 1 D j l i m t ) * = 1D j l L m t ) .
(- 1)m-ml (kLm (A.140)
To calculate the character X j ( g ) of a representation Dj it suffices to consider
a rotation about the z-axis, i.e. to set 6 and to zero. Rotations g with all
$J

other values of 6 and space $, but the same value of cp, belong to the same
class and thus have the same character X j ( g ) . Therefore

(A.141)

Sometimes the direct product of two irreducible representations Dj and Dji


is needed. It contains each of the representations Dl with values of 1 between
+
I j - j ' I and j j' exactly once. Thus the decomposition of the product
reads as

vj x vjt = c n.
j+jt

I= ( j - j l I
(A.142)

It is also worth mentioning that the generator matrices give rise to an irre-
ducible representation of the rotation group in yet another sense than that
discussed before: they are themselves the basis vectors of a representation.
The dimension of this representation is 3, i.e. the generator matrices trans-
form according to the vector representation of the rotation group. If one adds
inversion to this group, the result is the pseudovector representation. The
reason for this is that the sign of the generator matrices cannot change under
an inversion because of the commutation relations (A. 121). The transforma-
tion of the generator matrices as pseudovectors is employed in the invariant
method for construction of the k.p-Hamiltonian (see section 2.7). For that
purpose, the explicit form of these matrices is needed. By means of formu-
lae (A.136) and (A.137), one obtains, in the 2-dimensional case, the Pauli
spin-matrices multiplied by 1/2. For d = 3, it follows that
A.5. Representations of the full rotation group 68 1

(A.143)

( 0 0 -1
and for d = 4 the result is

0 id3 0

Id3
2 0 1
( 8 , 11, I &d)=
0 1 0

0 0 $d3

0 -$& 0

id3 0 -i

0 i o
0 0 id3
% o 0

0 ; 0
(A.144)
0 0 -4
0 0 0 - 32

Again, transformation of the pseudovector of three generator matrices re-


flects the close relation between these matrices and the angular momentum
operator J. This relation may be expressed as follows: The generator matri-
ces of dimension d are the matrices which represent the angular momentum
operator in the basis of the irreducible representation of the full orthogonal
group of this dimension. The basis itself may be characterized in this con-
text as the set of simultaneous eigenvectors of the square 5' of the angular
+
momentum operator J for eigenvalue 7i2j(j 1) with j = (d - 1)/2, and of
the z-component J z of J.
682 Appendix A. Group theory for applications in semiconductorphysics

A.6 Spinor representations

A .6.1 Space-dependent spinors


Having completed the discussion of irreducible representations of the full
rotation group in the previous section, which, as noted above, have only in-
direct significance for us, we now return to the representations of the finite
groups of rigid displacements which are of immediate interest here. The
smallest possible dimension of an irreducible representation of the full rota-
;
tion group is 2, corresponding to the smallest possible value of j . For the
pertinent representation matrix D i ( a ) , it follows from (A.138) that
2

(A'. 145)
The basis vectors -2 1 , - 2 - ~ which transform according to this representation,
5 2
are referred to as two-component spinors. We abbreviate them by Ii), i).
I-
In quantum mechanics, the states cp of electrons are described by just such
spinors with the two components (slcp) cp(s), s = i,
-$, corresponding to
the two possible spin states. The two spinor components are themselves vec-
tors in ordinary Hilbert space R having components (xlcp(s))z p(x,s) with
respect to the position vector x. For the quantum mechanical description of
electrons, the representations of the group of rigid displacements G in the
Hilbert space R of the two-component position dependent spinor functions
cp(x) are of central importance (the notation R should not be mixed with
that for a lattice vector). By definition, these representations imply that the
elements g of the displacement group G are uniquely assigned to a group of
linear operators g in the spinor Hilbert space R.Therefore, we must have

g1 . g2 = g 1 . g2. (A.146)
We now proceed to construct this assignment explicitly. To this end, we set

where S ( g ) denotes an operator in the 2-dimensional spin space. Since trans-


lations do not act on the spin coordinate s, for elements g = t, . a we have

S(g) S(t,. a ) = D I ( a ) . (A. 148)


2
For pure translation groups the only representation in spin space is the
identity representation. The group of rigid displacements can therefore be
A.6. Spinor representations 683

restricted to point and space groups. Using the basis I s) in spin space,
relation (A.147) can be written in the form

(A.149)
S

Since, according to the definition (A.46) of g, the function (o(g-lx,s) in


(A.149) equals 3 cp(x,s), and since the operator 9 can be factored out of the
sum, we have

The representations of the group G of rigid displacement in the spinor Hilbert


space R are thus the direct products of representations in the Hilbert space
R of scalar functions with the representation Vl of the full rotation group.
In the next subsection we therefore explore the ;epresentation V1 in greater
5
detail.

A.6.2 Representation D+
The two Euler angles cp and $ both vary between 0 and 2n. The ro-
+ +
tations a[cp 2a,8,$] and a[cp,8, $.J 2n] are identical with the rotation
a[cp,O,$]. Strangely enough this does not also hold for the pertinent ma-
+ +
trices D g ( a [ p 2n,8, $]) and D;(a[cp, 8, $ 2n]) from relation (A.145).
Both equal -Dr(a[cp, 8, $1). As long as one considers only single rotations
2
a [ p ,8, $1 this does not cause any difficulty because one can always restrict cp
and $ to the interval between 0 and 2n, and assign + D i ( a ) to the rotation
3
a. Difficulties arise, however, if two rotations a1 and a 2 are multiplied. Then
it may happen that one of the two angles (o12 or $12 of the product a1 . a 2
takes a value larger than 2n. In this case, the representation (A.145) assigns
the matrix - D ~ ( a 1. a 2 ) to a 1 . a 2 . As it is itself an element as well, a1 . a 2
2
has, however, already been assigned the matrix + D l ( a l . 4 ,which cannot
be changed because the assignment in a representation must be unique. The
only resolution of this difficulty is to allow the multiplication rule for the
two representation matrices D l ( a l ) , D i ( a 2 ) to differ from that of the two
group elements a 1 and a 2 by a minus sign. More precisely, the sign in this
rule must be allowed to depend on the particular combination of the two
elements a1 and a2. It must be a function w(a1, a 2 ) of a 1 and a 2 , with two
possible function values +1 and - 1. The corresponding multiplication rule
reads
684 Appendix A . Group theory for applications in semiconductor physics

It is obvious that the matrices Di(cu) obeying this multiplication rule no


longer form a representation of the group of rotations in the sense of section
A.4 according to the understanding there the factor w ( q , az) had to be +1
~

for all cu1, a 2 which is not the case here. In fact, assignments of matrices to
group elements obeying the multiplication rule (A. 151) are representations
in a generalized sense, that of so called spinor representations. They form
a special case of projective representations or representations with factor
systems which will be considered below in greater detail.
For the complete determination of the representations of point and space
groups in spinor space we still need the matrix D i ( I ) for the inversion I . In
determining it one can take advantage of the fact?that I commutes with all
elements g of the space group. Therefore D i ( I ) must also commute with all
D i ( g ) . It follows from a theorem of group theory (known as Schurs l e m m a )
2
that D i ( I ) must then be a multiple of the identity matrix 1. Because of
I
[ D I J I ) ]=~ 1, only the identity matrix itself or its negative are allowed. The
two assignments

D.(I) = +1 or D l ( 1 ) = -1 (A.152)
I
give rise to two inequivalent irreducible representations of point and space
groups containing the inversion.

A.6.3 Irreducible spinor representations


In the spinor Hilbert space R, the two spinor components p(s) are func-
tions p(x,s) of the position coordinate X. It has already been demonstrated
in subsection 1 that the representations of the displacement group G in R
are given by direct products of the representation V L of G and a particular
2
representation of G in the Hilbert space R of scalar functions. While Vi
2
signifies a spinor representation, the representations of G in the ordinary
Hilbert space R mean representations in the ordinary sense, often referred
to as vector representations. The direct product of a spinor representation
with a vector representation is again a spinor representation for the corre-
sponding multiplication rule is of the general form (A.151). From this we
may conclude that any representation of G in the spinor Hilbert space R
will not be an ordinary, but a spinor representation. On the other hand, the
direct product of two spinor representations yields a vector representation
because the multiplication rules for the product representation are of the
general form (A.59).
Of particular interest among the spinor representations of G are, of
course, the irreducible ones. We now set out to show how to determine
them explicitly.
A.6. Spinor representations 685

A.6.4 Double group method


The irreducible spinor representations of G may be obtained by means of the
irreducible vector representations of a fictional group, the so-called double
group G which has twice as many elements as G. The double group G is
defined as follows. To begin with, we introduce a group ?? of linear operators
in spinor space, which consists of all representation operators 9 . 0 1 ( a )where
T
j is an element of the representation operator group G of G, and D 1 ( a ) is
T
given by equation (A.145). The first half of the elements of g, which will
be denoted by &, are taken to be those with the Euler angles p and $
of a [ p ,0, $1 in D.(a) restricted to the interval between 0 and 27r, and the
2
second half, which will be denoted by g>, are those elements for which one
of the angles cp and $ lies in the interval between 27r and 47r, with the other
one in the interval between 0 and 27~as before. The operator group ?? thus
contains twice as many elements as G. The second half of the elements
can be obtained by multiplying all elements of the first half by a rotation
through a[27~, 0, 01 or a[O,0,27~].Let Q be the operator for a rotation through
a[27r,O,O] . It then follows from (A.145) that D + ( Q ) is the negative of the
identity matrix, -1. Therefore, D i ( Q ) . D L ( Q )= 1. According to equation
2
(A.145), the rotation a[O,0,27r] is represented by the same operator D i ( Q ) .
2
Since D $ Q ) acts only in spin space, one has D l ( Q ) x = x. Moreover, it
holds that D L ( Q ). j = g . D 1 p ( Q ) ,i.e. the element D 1 ( Q ) commutes with
F
all remaining elements of G. Finally, we reverse the assignment of group
elements and operators and assign the operator fj< to the displacement g,
and the operator g> to the displacement Q . g = g . Q.
The group of displacements g and Q g is the double group G we need for
the construction of the spinor representations of the simple group G. The
elements of G will be denoted by 9.The set G of elements g , generally, does
not form a subgroup of G because the multiplication of two elements g : and
g b of G may lead to an angle cp or $ between 27r and 4%. Analoguously, the
set Q . G = G . Q of elements Q . g or g . Q is not a subgroup of G. Whether
the product g$ . gb belongs to G or Q . G can easily be determined. The
square 1 of the inversion I , for example, is the identity element E , and the
square cr2 of a reflection equals Q. This follows from the relations Oh = I . C2
for Oh or uz,= I . U2 for uv (see equation (A.42)).
Representing the double group G in spinor Hilbert space R means
uniquely assigning to each element g a linear operator g in this space. This
assignment is determined by that for the simple group in equation (A.153),
according to

(A.153)
686 Appendix A . Group theory for applications in semiconductor physics

It may easily be demonstrated that the assignment (A.153) forms an or-


dinary or vector representation of G'. It also holds that each irreducible
representation space R, of the double group G' simultaneously forms a rep-
resentation space of the group G because the elements of G cannot map out
of R, if none of the elements of G' does so. Moreover, as a representation
space of G, R, cannot be reducible. Indeed, if R, were reducible, then there
would be at least two irreducible subspaces of R,, one denoted by R,l, and
another denoted by R,2 which forms the complement of R,l in R,. The
two subspaces Rvl and Rv2 would be transformed only into themselves by
all elements g of G. However, R,l and R,2 would also be mapped into
themselves by all elements Q . g , because the application of Q results solely
in multiplication by -1. Thus R, would also be reducible as a representa-
tion space of GI, which contradicts the initial assumption. Each irreducible
vector representation of G', therefore, is also an irreducible representation
of G.
The latter may be either a vector representation or a spinor representa-
tion, both cases must be taken into account. Since one can also prove the
converse, i.e. that each of the irreducible vector or spinor representations
of G gives rise to a vector representation of the double group GI, it follows
that the set of irreducible vector and spinor representations of the simple
displacement group G is identical to the set of irreducible vector represen-
tations of the associated double group GI. Here we are only interested in
the spinor representations of the original displacement group G. These can
easily be selected from the whole set of irreducible vector representations
of G' - they are all those which are not also vector representations of G.
These irreducible vector representations of the double group are called eztra
representations. Using this term one may state the following theorem:

Theorem 11

The irreducible spinor representations of a displacement group G (point or


space group) are the extra representations of its double group GI.

The decomposition of double groups into classes is of importance for the


explicit construction of the irreducible representations of these groups. Like
the identity element E , Q also forms its own class. The reason for this is
the commutativity of Q with all elements g and Q . g . From this, it follows
that elements g which are conjugate in G are also conjugate in G'. The same
holds for the pertinent elements Q . g , also they are in the same class of the
double group G' if the elements g are in the same same class of the simple
group G. If each element Q . g fell into a class different horn that of g , then
the double group G' would always have twice as many classes as the simple
group G. However, this is not the case. Indeed, if a rotation axis of the
A. 7. Projective representations 687

order n is twesided (see section A.2), the elements Ck and CE-k = Q . Cqk
belong to the same class, and so do the elements Q . Ck and CGk, however,
the two classes are different. If an n-fold rotation axis is not two-sided, then
C;, C i k ,Q . Ck and Q . CLk form four different classes. For n = 2 the two
elements C2 and C,' of the simple group G are identical, but in the double
group G t they differ for it holds C; = Q which results in C2 = Q . C,' and
C', = Q . C2. The two elements C2 and C,' 3 Q . C2 form one class if
and only if the 2-fold axis is two-sided. An analogous statement holds for
reflections uh. In this case uh and Q . uh belong to one class if the 2-fold
rotation axis C2, associated with ah by the relation uh = I ' C2, is two-sided.
Similarly, u,, and Q . cr, are in one class if the rotation axis U2, associated
with u, by the relation u,,= I . U2, is two-sided.
We illustrate the above results using the tetrahedral group Td as an exam-
ple. In section A.3 we found 5 classes of the simple group T d : E , 8C3,6S4,3C;
and 60,. From them the following classes of the double group T d are obtained
by means of the rules mentioned above: ( E } , {Q}, {4C3,4Q . Cj}, (4Q .
C3,4C$}, {3C4,3Q.C,3},{3Q.C4,3c,3}, {3c,2,3Q.C,2}, { 6 u v 7 6 Q . ~ v }The
.
48 elements of the double group Ti belong therefore to 8 different classes.
Since the number of classes equals the number of different irreducible rep-
resentations, and since the simple group T d has 5 classes, there must be 3
irreducible spinor representations of the group T d . The sum of squares of
the dimension of these representations must be 24. There is only one way to
obtain the number 24 by summing three squares, namely 2' + +2' 4' = 24.
Thus two of the three spinor representations of Td have the dimension 2, and
one has the dimension 4.
The spinor representations of point and space groups were characterized
above as projective representations with a particular factor system. Later,
we will see that projective representations will also occur in another context,
namely the ordinary vector representations of space groups. These represen-
tations may be traced back to the projective representations of point groups
of equivalent directions with a particular factor system determined by the
crystal structure. This demonstrates that projective representations play
an important role in the applications of group theory in solid state physics.
In the following subsection, we discuss properties of these representations
which will be of use later.

A.7 Projective representations

A.7.1 Factor systems


Representations of groups are unique assignments of their elements to ele-
ments of groups of linear operators in a particular space. In regard to the
688 Appendix A . Group theory for applications in semiconductor physics

representations defined in section A.4, the multiplication rules of the op-


erator group could be derived from that of the original group because the
assignment was given a priori. According to these rules, the product of two
operators assigned to two group elements was equal to the operator assigned
to the product of the two elements. In the projective representations to be
considered below one takes the opposite approach: Not that the operators
ij assigned to the group element g are specified a priori, but rather the mul-
tiplication rules of the assigned operator group G are defined at the outset.
The assigned operators are later derived from these rules. The multiplica-
tion rules of the operator group are defined as follows. Let gl and 32 be
the two operators assigned, respectively, to the two group elements g1,gz.
Then complex numbers w(g1,gz) depending on g1,g2 are introduced and the
product 31 . 92 is defined as

91.92= w(g1,92) 31 92 . (A.154)

The set of factors w(g1, 92) is called the factor system of the representation.
The assignment of G to the operator group G defined by the multiplication
rules (A.154) is referred to as a projective representation or representation
with factor system.
The factors w(gI,g2) are restricted by the group properties of the oper-
ator group G. The associative property results in the condition

It may be shown that this condition is also sufficient to assure that G is a


group, i.e. to make G a projective representation of G.
If the absolute value of the factor system w(g1,gz) is 1, i.e. if

holds, then the corresponding projective representation of the displacement


group is unitary, as in the case of ordinary representations.
The ordinary representations are projective representations with the par-
ticular factor systemwt(gl,g2) = 1for an elements g1,g2. We callw!(g1,g2) =
1 the identity system. If a particular factor system w(gl,g2) is given, an infi-
nite number of other factor systems w(gl,g2) can be constructed by means
of an arbitrary complex function u(g), namely by setting

(A.15 7)
A. 7. Projective representations 689

This can easily be verified by demonstrating that equation (A.155) holds


for w(g1, g 2 ) if it holds for w(g1,92). Two projective representations D and
D whose factor systems w(gl,g2) and w ( g l , g 2 ) are related by the equation
(A.157), are called pequiwalent. The representation operators $ ( g ) and $ ( g )
assigned to the group element g in the two p-equivalent representations obey
the equation

(A.158)

The entirety of all p-equivalent factor systems is referred to as a class of


factor systems. If the factor system has the particular form

(A.159)

then it is p-equivalent to the identity system w t ( g 1 , g a ) = 1, and the cor-


responding class is called identity class. In this case the projective repre-
sentations are p-equivalent to representations in the ordinary sense. The
number of different classes of factor systems of a finite group is also finite.
The numbers and particular forms of these classes are characteristic for each
group. The identity class occurs for all groups. For cyclic groups, it is the
only one possible. Generally, the number of different factor system classes
is a power of 2 , therefore it is either 2O, or 2 l , or 22 etc. Most of the 32
point groups of crystal structures have several classes of factor systems, for
example, D4h has 8, Td has 2, and O h has 4. If there are only 2 factor sys-
tem classes, then the second class beside the unit class necessarily includes
the factor system of spinor representations because the spinor representa-
tions also always exist. In special cases (in particular, for the point groups
c1,c 2 , c 3 , c4,c6,c,,s4, C3h, D3, c3,,),the spinor representations coincide
with the vector representations. Then only the identity class of factor sys-
tems exists and the factor system of the spinor representations is p-equivalent
to the identity system.

A.7.2 Definitions and theorems

Unitary equivalence

Let us consider two projective representations 2, and V M with the same


factor system, and let R, and R,M be the corresponding representation
spaces. The representations two are said to be equivalent to each other if, as
in the case of ordinary representations, a unitary operator M exists which
transforms the two representation spaces R, and R,M into each other. To
690 Appendix A. Group theory for applications in semiconductor physics

avoid confusion with the term p-equivalence, one speaks of unitary equiv-
alence rather than just of equivalence, as we did in the case of ordinary
representations.

Reducibility a n d irreducibility

The concepts of reducibility and irreducibility can be defined for projective


representations as well. This can be done in the same way as for ordinary
representations, and we need not repeat the definition here. It also holds
that any unitary projective representation with a particular factor system
may be decomposed into irreducible projective representations having the
same factor system as the original projective representation.

Characters

The operators g representing a group element g in a particular projective


representation may be written as matrices D ( g ) with respect to a particular
basis of the representation space R,. As for ordinary representations, the
character X ( g ) of an element g is defined as the sum of the diagonal elements
of its representation matrix D ( g ) . Also, the characters of unitarily equivalent
projective representations are identical. However, unlike the case of ordinary
representations, for projective representations the character of an element is
no longer only a function of its class. As a consequence of this, the number
of unitarily inequivalent irreducible projective representations of a group
G no longer equals the number of classes of conjugate elements G, but is
smaller than the latter one, unless the factor system class of the projective
representation is that of the identity system. In the case of the irreducible
spinor representations of the point group T d considered above, we already
noted that there are 5 classes of conjugate elements in T d , but only three
spinor representations.
If one knows the character X ( g ) of a projective representation, then the
character X(g) of a p-equivalent representation defined by equation (A.157)
is given by the relation

X(g) = 4 7 ) x ( g ) . (A ,160)

If the matrices D ( 9 ) form an &dimensional irreducible projective represen-


tation with a particular factor system, then it follows from (A.160) that the
matrices D(g) = u ( g ) D ( g ) form an irreducible projective representation of
the same dimension d with the pequivalent factor system given by equation
(A. 15 7).
Regarding the dimensions of irreducible projective representations of
groups G possessing more factor system classes than just the unit class, one
A. 7. Projective representations 69 1

may show that the representations for factor system classes different from
the identity class must be 2- or higher dimensional. That this rule works
may be seen from the irreducible representations of the group T d considered
in the section on spinor representations. This group has two classes of factor
systems and, therefore, one set of non-ordinary representations. These are
the spinor representations of T d , and none of them is 1-dimensional.

Orthogonality

The orthogonality relations (A.71), (A.72) and (A.81) for the matrix el-
ements and characters of ordinary irreducible representations also remain
valid for projective irreducible representations with the same factor system.
Matrix elements and characters of irreducible representations with differ-
ent factor systems are always orthogonal to each other. Owing to these
properties of the characters, the same procedure for the decomposition of a
projective representation into irreducible parts can be applied as that used
for ordinary representations, namely, in accordance with equation (A.83),
the projection of the character of the representation onto the characters of
the irreducible representations. Being a consequence of the orthogonality
relations, Burnsides theorem holds also for projective representations:

Burnside theorem

The sum of the squares of the dimensions of all unitarily inequivalent irre-
ducible projective representations of a group having the same factor system
equals the order of the group.

From this theorem and the statement on the dimensions of p-equivalent


representations in connection with equation (A.160) it follows that for each
projective irreducible representation with a particular factor system, there is
exactly one projective irreducible representation with a p-equivalent factor
system having the same dimension. If one, therefore, knows the projec-
tive irreducible representations for one factor system of a class, one also
knows these representations for all the other factor systems in the same
class. The projective irreducible representations with factor systems from
different classes will, however, differ in general from each other regarding
their dimensions and numbers.

Direct product

The direct product of two projective representations is defined as it was for


ordinary representations, i.e. as the representation induced in the product of
the representation spaces. The factor system of the product representation
equals the product of the factor systems of the two projective representations
692 Appendix A . Group theory for applications in semiconductor physics

which are multiplied. A particular consequence of this rule is that the factor
system of the product of two spinor representations for which the factors
are either 1 or -1, equals the identity system. This confirms our former
conclusion that the product of two spinor representations forms an ordinary
or vector representation.
The consequences of spatial symmetry for the matrix elements of opera-
tors as expressed in theorem 10 are also valid if the wavefunctions transform
according to projective and, in particular, to spinor representations.
Finally, we sketch how irreducible projective representations may be con-
st ruct ed.

A.7.3 Construction of projective representations


The irreducible projective representations may be constructed in a manner
similar to the construction of the spinor representations in section A.6. Us-
ing the particular factor system for spinors defined by equations (A.145) and
(A.151) a double group G was introduced in addition to the original group
G. The irreducible vector representations of the double group resulted in
two sets of irreducible representations of the group G , namely the vector
and spinor representations. It has already been pointed out that the lat-
ter are projective representations of G with the particular factor system of
spinors. The above procedure may be generalized to obtain the irreducible
projective representations for all possible classes of factor systems. In fact,
this procedure was originally derived in the general case of projective repre-
sentations by Schur, and later Bethe applied it to representations in spinor
space. Based on the various classes of factor systems of G , Schur constructed
a super-group G of as many times more elements as there are different factor
system classes. He then demonstrated that the ordinary irreducible represen-
tations of this super-group G yield all irreducible projective representations
of all classes of factor systems of the original group G . The explicit construc-
tion of the super-group which delivers all the possible factor system classes
exceeds the limitations of this introduction to group theory. The reader can
find it in the book by Bir and Pikus (1974). This book gives also the irre-
ducible projective representations of the 32 point groups of crystals for all
of their factor system classes.

A.8 Time reversal symmetry


The state of an electron generally depends on time t , i.e. its components cp(x)
in ordinary Hilbert space or cp(x,s ) in spinor Hilbert space are, respectively,
functions cp(x,t ) or cp(x,s,t ) of t. As well as transformations in coordinate
and spin spaces, one can also consider transformations on the time axis.
+
Examples are translations of time by a constant to, i.e. t becomes t to, and
A.8. Time reversal symmetry 693

time inversion i.e. the transition from t to -t. The latter transformation is
commonly referred to as t i m e reversal, and is denoted by the symbol K .

A.8.1 Time reversal operator

As in classical mechanics, the basic dynamical laws of quantum mechanics


are invariant with respect to time reversal. In short, we say that they have
t i m e r e v e r s a l s y m m e t r y . This symmetry holds, in particular, for the time-
dependent Schrodinger equation governing the time development of electron
states lp(x,t ) without spin, as well as for the Pauli equation which applies
to electron states lp(x,s, t ) with spin. From this observation one can deduce
the transformation of the wavefunctions under time reversal. If the time t
is reversed in the Schrodinger equation, then the vector lp must be replaced
by the complex conjugate vector lp* in order that the equation remain valid.
The time reversal transformation K therefore causes the transformation of
lp to lp*. In other words, in the Hilbert space of scalar functions the time
reversal operation K is given by the operator K defined by the relation

Klp = lp*. (A.161)

If there is a magnetic field, one also has to reverse its direction in order that
lp* satisfy the Schrodinger equation.
In order that the Pauli equation remain valid under time reversal, the
operation K must be represented by an operator K in spinor Hilbert space
defined by

being the y-component of Paulis spin matrices. One can easily show that
for spinors the relation (klp I lp) = 0 holds, i.e. Klp is orthogonal to lp. If
lp is an eigenstate of the Pauli Hamiltonian for a given energy, and if there
is no magnetic field, then Klp forms a second linearly independent eigen-
state for this energy. That means that, because of time reversal symmetry,
all eigenvalues of the Pauli equation are at least doubly degenerate in the
absence of a magnetic field.
Without spin it is evident that K 2 = E , and K 2 = -E with spin.
This implies that without spin the unit element E and the time reversal
operation K form a group, while with spin the elements E , K , K 2 and K 3
do so. We denote both groups by 2 . Since K commutes with all elements
of the group G, the direct product G x 2 of G and Z is itself also a group.
We call it the space-time g r o u p and denote it by 6 . We are interested in the
694 Appendix A. Group theory for applications in semiconductor physics

joint effect of spatial and temporal symmetry of a system. At first glance


one could think of exploring these by using the irreducible representations
of the space-time group Q defined above. This is, however, not possible,
because of the unusual properties of the time reversal operator K . This
+
operator is not linear, since one has %[clcpl c2p2] = [crkpl czkpn], +
i.e. one gets the complex conjugate coefficients c?,cf while the coefficients
themselves would be needed for f? to be linear. The operator K is also
not unitary, since one has (Ifpl 1 Ifpz) = (pl I p2)*, while (91 I p 2 )
should follow for I? t o be unitary. In the group representations considered
so far, the linearity of the group operations played a fundamental role. In
particular important it was for theorem 1, which states that a subspace
of eigenfunctions of the one-electron Hamiltonian for a particular energy
eigenvalue is also a representation space for the full symmetry group of the
Hamiltonian. This theorem does no longer hold if time reversal is included,
i.e. if the space-time group 4 is taken to be the symmetry group instead
of the space group G alone. For these reasons, the powerful tools of group
representations, which apply in the case of space symmetry, are no longer
available if the joint consequences of space and time symmetry are to be
studied.
Mainly two questions need to be addressed, namely, whether time re-
versal symmetry of the Hamiltonian causes additional degeneracy of energy
eigenvalues, and secondly, what additional relations emerge among the ma-
trix elements of operators if time reversal symmetry is taken into account.
These two questions will now be directly answered.

A.8.2 Additional degeneracies of energy eigenvalues

We start by solving the first problem. To this end we consider an irre-


ducible representation 2) of G in the Hilbert space of scalar functions p(x)
or of spinors p(x, s) in spinor Hilbert space - both possibilities are admit-
ted below. Let 2) be a vector representation or a spinor representation,
as needed. We denote the representation space in both cases by R D , and
a basis therein by 91,p 2 , . . . , v d . The vectors Kpl,Kpa,. . . , K p d span a
space R p which is likewise an irreducible representation space of G for K
commutes with all elements of G. The representation of G in RD* is the
complex conjugate representation D* of D. This follows from the relation
(kpi I g I K P k ) = (cpi I g 1 p k ) * for the matrices of the two representations,
which holds both with and without spin. If the two spaces R D and RD* are
joined, a representation space RD of G emerges which is generally reducible.
The dimension of any irreducible representation subspace in RD cannot be
smaller than d , or else D and D* are reducible. The dimension 6 of RD can
be at most 2d. The possibility that 6 lies between d and 2d has to be ex-
A.8. Time reversal symmetry 695

cluded, however, because then an irreducible representation space of smaller


dimension than d would exist, namely the complementary subspace of R D in
RD.Under G , this subspace would transform into itself as RD and R D do.
It would, therefore, form an irreducible representation space of dimension
d - d < d in RD,which is impossible. It follows that the dimension of 720
will be either d or 2d. An equivalent formulation of this result is that the
basis functions Kpl, Kp2,. . . ,K q J d of RD* are either all linearly dependent
on the basis functions p1,p2,.. . , p d of R D , or all linearly independent of
them.
If G is understood as the spatial symmetry group of the crystal Hamilto-
nian H , and if in addition there is time reversal symmetry, the two basis sets
p1, p 2 , . . . , pd and Kpl, K p z , . . . , K p d together can be identified with the
totality of eigenfunctions of H for a particular energy eigenvalue ED. The
degeneracy of E D is d if the two basis sets are linearly dependent on each
other, and 2d, if they are linearly independent. The former case is referred
to as case a). In the latter case a further distinction is necessary. The rep-
resentations V and V* which belong, respectively, to the two subspaces RD
and R p of 7 2 ~are , either inequivalent, which will be case b), or equivalent,
which will be case c). In case a ) the degeneracy of an energy eigenvalue
does not change because of time reversal symmetry. Using the fact that the
two equivalent representations V and V * are representations in the same
space R D , one can show that a real representation exists in R D which is
equivalent to V and V*. In cases b) and c), the degeneracies double b e
cause of time reversal symmetry, i.e. they are twice the degeneracies due to
spatial symmetry alone. The doubling occurs in both cases because two dif-
ferent irreducible representation spaces of the space group are joined to one
eigenspace of the Hamiltonian because of time reversal. In case b), the two
joined representations are inequivalent, and in case c) they are equivalent.
We can also say that in case b) the characters of the joined representations
are both complex and one is the complex conjugate of the other. In case c)
the two joined representations have the same characters, and the characters
are real. This does not mean, however, that the representations are them-
selves real. They are complex, in fact, unlike case a) where they are real.
The relationships described above are summarized in Table A.3. They are
commonly referred to as the Herring criterion.
The question of which of the three cases a), b) or c) applies to a particular
representation V ,is answered by a theorem which establishes a connection
between the above three cases and the sum C X ( g 2 ) of the characters of all
squared elements g2 of G.

Theorem of Frobenius and Schur

This theorem is expressed by the relation


696 Appendix A . Group theory for applications in semiconductorphysics

Table A.3: Properties of the time-reversed irreducible representations of space


groups and consequences of time reversal symmetry for additional degeneracies of
energy eigenvalues.

Herring Relationship Relationship of V Degeneracy


case of R D and Rb and D*

RD= Rb l? equivalent to V * Unchanged


and both equivalent to
real representation
Xb = X D

RD# R b 2) inequivalent to V* Doubled


Xbf XD

RD# Rb 2) equivalent to I?* Doubled


but both not equivalent to
real representation
Xb = X D

K~ in case a)
1
- c
gEG
X(g2)= 0 in case b) (A.163)
- K ~ in case c),

where K 2 = 1 for vector representations, and K 2 = -1 for spinor repre-


sentations. We note that cases a) and c) are interchanged if one switches
over from an ordinary to a spinor representation. The value +lon the
right hand side of equation (A.163) means no additional degeneracy for vec-
tor representations, but an additional degeneracy for spinor representations,
while the value -1means additional degeneracy for vector representations,
but no additional degeneracy for spinor representations. The definitions of
cases a) and c) were originally made using the values +land -1 rather
than K 2 and - K 2 as we do here (for good reasons). One has to keep this
in mind to avoid confusion.
A.8. Time reversal symmetry 697

A.8.3 Additional selection rules for matrix elements


Time reversal symmetry has particularly simple implications for matrix el-
ements of operators A(t) in Hilbert space which are either even or odd
functions of time, such that

KA(t)K-= nA(t) with n= 1 or n = -1. (A.164)


As in section A.4, we consider matrix elements (p& 1 A k ( t ) 1 cpr)of tensor
components A k ( t ) of operators A(t) between the basis vectors {py, cpz, . . .,
~p:~} and {cpy, p$, . . . , of particular irreducible representations V, and
Dp of the space group G. For the matrix elements of &(t) between the time-
reversed basis vectors {Kpy, Kp!j, . . . ,Kcpzp} and { K c p y , Kpt, . . ., K p z U }
it follows

(G4I A&) I Kcp3 = 4 P k I A k ( t ) I cpY)*. (A.165)


This relation implies that the matrix elements with respect to timereversed
basis vectors are completely determined by the original elements involving
the non time-reversed basis vectors. If either or both of the two represen-
tations D, and V p belongs to case a) defined above , then { K c p y , Kp;, . . .,
KpzY} are linearly dependent on {cpy, pz, p$} or {Kpy, Kcp!,. . . ,K p S p }
are linearly dependent on py, pg, . . . , p k or there is such a linear depen-
dence for both. Therefore equation (A.165) leads to relations between the
elements (cpc cpr)
1 Ak(t) I themselves and the number of independent matrix
elements (qkI A k ( t ) 1 pr) is reduced by time reversal symmetry. Below, we
will outline a procedure which allows one to calculate the reduced number
o i independent elements in the special case p = v.
If neither of the representations V v and Dp belongs to case a), then
equation (A.165) does not impose an additional condition on the matrix
elements (cp& I A k ( t ) I pr), rather it connects these elements with matrix
elements between basis vectors of other representation spaces.
Consider now the mixed matrix elements (Kcp2 I A k ( t ) I pr) and (cpk I
A k ( t ) I I?&) between one original basis vector and a time-reversed one. For
these, one can easily derive the relations

(& I I KPK) = .K2(cpF I Am I &4J . (A.167)


These equations mean that the number of independent matrix elements
( g p & ( A k ( t ) 1 pr)?as well as the number of independent matrix elements
(pg 1 A&) 1 Kply))is reduced because of time reversal symmetry. The
factors n and K 2 in (A.166) and (A.167) can take, independently of each
698 Appendix A . Group theory for applications in semiconductorphysics

other, the values +1 or -1, so that the product K K will ~ be either +1 or -1.
The matrices (I?& I Ak(t) I pr) and (pg I A k ( t ) I Kpr) are thus either
symmetric (nX2 = +1) or antisymmetric ( & K 2= -1) with respect to the
exchange of the pairs of indices p m and vl. How the number of independent
matrix elements follows from this observation in the general case is shown
in the book by Bir and Pikus (1974).
In the special case p = v , one can apply the procedure developed in
section A.4.3 for the determination of independent matrix elements without
taking account of time reversal symmetry, provided one makes a distinction
between symmetric products [D,x D,],and antisymmetric products [D,x
D,Ia, as defined in section A.4.3. The matrix elements (Rp; 1 A k ( t ) I p?)
transform themselves according to the representation [a,x v,]~ x Dd, if
nK2 = +1, and according to the representation [D,x V,Ia x V A ,if K K =~
-1. Following section A.4.3 the matrix (zpk I A ( t ) I (or)then has as many
independent elements as the number of times the identity representation
occurs in the product representation [V,x D,],x V A for K K =~ +l, or
in the product representation [D,x V,la x D A for r;K2 = -1. Analogous
statements hold for the matrix (p; I A&) I kp?), where V, is to be
replaced by VE.

A.9 Irreducible representations of space groups


Below, we discuss the irreducible representations of the space groups in the
Hilbert space of one-electron states p(x). Initially, electron spin will be
omitted, i.e. we start with the ordinary or vector representations. As out-
lined in section A.4.1 the wavefunctions p(x) are taken to be periodic with
respect to a periodicity region of the crystal, so that the originally infinite
space groups become finite. The aim of the present section is to develop
a procedure to derive the general form of the irreducible representations of
space groups. Using this procedure, the representations of two particular
space groups, those of diamond and zincblende crystal structures, will be
obtained in section A.lO.

A.9.1 Representations of translation groups


Each irreducible representation of a space group G is also necessarily a rep-
resentation of the translation group T contained in it. As such, it is in
general reducible. Since T is an Abelian group, the irreducible represen-
tations of T are 1-dimensional. This means that the representation space
of an irreducible representation of a space group G can be set up from 1-
dimensional irreducible representation spaces of the translation subgroup T
of G. We mark these spaces by a vector k whose meaning will immediately
become clear. The spaces themselves, which are 1-dimensional as we know,
A.9. Irreducible representations of space groups 699

are denoted by ( p k . The irreducible 1 x 1 representation matrices of the G3


elements t~ of T with respect to (pk are complex numbers c k ( R ) , given by
the relation

-ik.R
ck(R) = e (A. 168)
as was demonstrated in Chapter 2 of the main text. In this, k i s a vector in
a vector space reciprocal to the space of position vectors x . We have

k = kibi + k2b2 + k3b3 , (A.169)


with bl, b2, b 3 being primitive lattice vectors of the reciprocal lattice and
k l , k2, k3 being real numbers. The periodicity condition with respect to
the periodicity region causes the components of k to be of the form k l =
(2n/G)ll,k2 = (27r/G)l2, k3 = (2n/G)13with 11,12,/3 as integers. If K is an
arbitrary vector of the reciprocal lattice, then the irreducible representation
v k of the translation group T belonging to k is identical with the irreducible
representation V k f K belonging to k+K, since e x p ( 4 K . R ) = 1. One there-
fore gets all different irreducible representation of the translation group if k
varies within a primitive unit cell of the reciprocal lattice. Altogether, there
are G3 such vectors and thus also G3 different irreducible representations of
the translation group. This number also follows from the Burnside theorem,
- since the irreducible representations are 1-dimensional, there are as many

inequivalent irreducible representations as there are group elements, i.e. G3.


By definition, the effect of the translation operator t R on the basis function
vk(x) of the irreducible representations is given by

(A.170)
The basis functions (pk(x) may be thought of as Bloch functions, i.e. as
plane waves modulated by a lattice periodic factor u ~ ( x ) ,

(A.171)

Bloch functions (pk(x)of the same wavevector k, but different Bloch factors
Uk(x), give rise to the same representation of the translation group. As long
as one considers only the translation group, no relations exist between the
Bloch factors of the basis functions p k for different values of k - any Bloch
function can be used for a given k. Below, starting from an arbitrarily chosen
(but later h e d ) Bloch function ( p k ( x ) of wavevector k,we will construct basis
functions for an irreducible representation of the space group G. It turns
out that basis functions cpk,(x) will emerge in this process which transform
themselves according to irreducible representations of the translation group
with wavevectors k' # k. The Bloch factors up(x) of the basis functions
700 Appendix A . Group theory for applications in semiconductor physics

~pk~(x) of these representations cannot, however, be chosen arbitrarily, but


are determined by the Bloch factor U k of ' p k . The determination of the
wavevectors k' and the pertinent Bloch factors Uk' is the essential task which
has to be solved in constructing the irreducible representations of a space
group G from those of its translation subgroup T . The primitive unit cell in
which the wavevector k can vary will be chosen below to be the Wigner-Seitz
cell of the reciprocal lattice, i.e. the first B Z . The reason for this is that the
latter cell, generally, is the only primitive unit cell of the reciprocal lattice
which exhibits the full symmetry of the point group of equivalent directions.

A.9.2 Star of wavevectors


According to section 1, each element g of the space group G may be written
as product g = t ~tT(a). ' a of an orthogonal transformation a , a translation
tT(.) by a fractional lattice vector ? ( a ) associated with a , and a lattice
translation t R belonging to the translation subgroup T of G. The orthogonal
transformations a form the point group of equivalent directions P of the
crystal. The effect of an element g of the space group on 'pk(x) is described
by

3 rPk(X) = Pk(g-lx) = ,-iak.R cPlc(a-l[x - 5(a)1). (A.172)


Here, a k denotes the vector to which k transforms if the orthogonal trans-
formation a which originally acted on the lattice vector R, is transferred to
the vectors k of reciprocal space, as was explained in Chapter 2 of the main
text. Rewriting equation (A.172) in the form

it becomes clear that [tT(a) . a] 'pk(x)is a function which transforms accord-


ing to an irreducible representation of the translation group with wavevector
ak. The set of vectors ak generated as a runs through all elements of the
point group of equivalent directions P , is called the star {k} of k.
Henceforth, we must distinguish between vectors k with and without
symmetry. In the case of non-symmetric k, we have ak # k K for all p +
elements a of the point group P , where K is a vector of the reciprocal lattice.
For symmetric k we have a k = k+K for at least one element (Y different from
the identity element of P . Under the action of such an element, the function
[tT(a). a ] ( p k ( x ) transforms according to the same irreducible representation
of the translation group as does the function 'pk(x).
We consider, initially, non-symmetric k. The star {k} has p different
points crlk 3 k = k1,agk = k2,. . . , kp = q,k, where a j , j = 1 , 2 , . . . , p ,
denotes the elements of the point group P of equivalent directions. With
A.9. Irreducible representations of space groups 701

each star point kj, there is an associated basis function (Pg. We define the
latter by the relation

(A.174)
For the Bloch factor Uki(x) of ' p k j ( x ) , one then has

One can easily show that the space R { k } spanned by the p functions ' p k j ( x )
of equation (A.174), forms a representation space for the space group G. To
prove this, we write the elements t R . t T ( a j t ). a y of the space group G in the
form

z= kT(aj,) . aj'l ' tajrlR. kT(oIj) ' a j l -1 I (A.176)

where j', at fixed j , takes all values between 1 and p . That (A.176) is indeed
a possible way of writing of the space group elements g can be seen as follows:
first one commutes the translation t R in t R . t T ( a . , ). ajt with [tT(a.,). ajt],
simultaneously replacing t R by tujt-1 R in accordance with equation (A.17).
Multiplying [tT(ajt). aj,]. t ai-1l R with the particular space group element
[t,(,,, . aj]-', one obtains again an element of the space group. If j' runs
through all values between 1 and p , and R through all lattice points, then all
elements will be different, in accordance with the general group properties
derived in section 1. The set of these elements coincides, therefore, with the
space group G. This proves that the space group elements can indeed be
written in the form (A.176).
The application of g in the form (A.176) to one of the basis functions
f , D k j ( x ) of R { k } of (A.174) yields, in view of (A.173), the expression

(A. 177)
The function f,Dk,,(x) and therefore also g'pk2 ( x ) lie in R { k } . This means
that R { k } forms a representation space. R { k } is also irreducible, we omit
formal proof here. For non-symmetric k, the irreducible representations of
the space group can therefore be characterized solely by the k-vector itself
or, more strictly, by the star {k} of this vector. Thus, in this context we
write Z ) { k ] for the irreducible representations. The character x { k ) ( g ) of an
element g in the representation Z ) { k } reads

P
x { k } (9) = e- ikj-R- (A. 178)
j=1
702 Appendix A . Group theory for applications in semiconductor physics

A.9.3 Small point groups and their projective representa-


tions
Now let k be a symmetric point. The elements Q of P which transform k
into itself or into an equivalent vector k + K , form a subgroup Pk of the point
group P . It is called the small point group ofk. The corresponding elements
[tR.t,(,) . a ]of the space group form a subgroup G k of the space group G. It
is called the small space group of k. The latter subgroup Gk of G plays a role
similar to that of the translation subgroup T in the case of non-symmetric
k-vectors. Unlike T , Gk is not an Abelian group; thus the irreducible repre-
sentations of Gk, unlike those of T , are not necessarily 1-dimensional. This
also implies that there may be more than one irreducible representation Dk
for a given k (later we will distinguish these representations by an additional
index v, here we keep the general notation D k ) .
w e assume that a particular representation Dk of Gk is known. The
pertinent representation space will be denoted by Rk, and a basis therein by
p k l , p k 2 , . . . ,p k d . According to the assumption, for an element g of G k , the
function

gpkn(X) [ t R . t7(cr) . a ] V k n ( X ) = ( o h ( a - l [ X - .(a) - R]) (A.179)

also lies in Rk. Therefore, it may be written as

3 (Pln(x)= ,-ik.[?(u)+R] C~mn(Q)Vh(X) (A.180)


m

where Dmn(a) denote the expansion coefficients. This relation means that
the assumed representation of the small space group assigns to each element
Q of Pk a specific matrix Dmn(a). Below, we will demonstrate that the

matrices Dmn(a) form a group with multiplication rules of the general form
(A.154), and thus, a projective representation of the small point group 9.
To this end we consider two elements g 1 = [tR1 . t7(,l) . 311 and 9 2 = [tRz
t T ( a 2 ) az] of Gk. For brevity we set T ( Q ~ =
) 7-1, and 7 - ( ~ 2=
) 7-2. Equation
(A.180) associates the elements g1 and 9 2 , respectively, with the matrices
Dmn(Qi1)and Dmn(~12). For the effect of the product element [tR1. t , . a l l .
[ t R z . t, . a21 on pkn, one obtains from (A.179) the relation

(A.181)
mmf

On the other hand, using (A.45), we also have


A.9. Irreducible representations of space groups 703

(A.182)

- ,-wRl+fi+al(R2+ii)I CDm,n(al
. a2)cPkm,(y+ (A.183)
m'

Comparing equations (A.181) and (A.183), one finds that the matrices D(a1),
D ( a 2 ) and D(a1 . a 2 ) are related by the equation

where

Hence, the assertion is proven. The matrices D ( a ) do indeed form a pro-


jective representation of the small point group Pk of k. The factor system
is given by equation (A.185). The projective representation of the small
point group Pk arises from the representation of the small space group Gk
assumed at the outset. If the assumed representation Z)k of Gk is irreducible,
the same is true of the pertinent projective representation of Pk. Vice versa,
an irreducible projective representation of PI, gives rise to an irreducible
representation Z)k of Gk.
The factors w ( q , a2) of (A.185) will be referred to as the crystallographic
factor system. We will rewrite these factors in a more convenient form.
+
The scalar product k . al(R2 ?2) in the exponent may be expressed as
-1
a1 k.(R2+7'2). The exponent in (A.185) then becomes i(k-a11k).(R2+?2).
Here, a1 is an element of the small point group of k, i.e. q k may differ
from k by a reciprocal lattice vector K only. Because exp(-zK . Rz) = 1,
we may omit R2 from the exponent, whence

(A. 186)
For symmorphic space groups, no fractional translations 7' occur, i.e. the
vectors ?(a) are zero for all elements a of the small point group of a sym-
metric k-vector. Thus, the crystallographic factor system equals the identity
system for all symmetric k-vectors, and the projective representations of the
small point groups become representations in the ordinary sense. For non-
symmorphic space groups, a factor system different from the identity system
may occur, but only for vectors k on the surface of the first B Z . In fact,
if k is an internal point of the first B Z , then ak also lies in the interior of
704 Appendix A. Group theory for applications in semiconductorphysics

the first B Z and cannot differ from k by a non-zero reciprocal lattice vector
K. For surface points, however, it is possible that ak deviates from k by a
vector K # 0. If that is the case, then factors differing from 1 may occur
in the crystallographic factor system (A.185). The basis ( p k l , ( p k 2 , . . . , (pkd of
the irreducible representation v k of the small space group G k then gives rise
to a projective irreducible representation of the small point group P k of k.

A.9.4 Representations of the full space group


For any irreducible representation of the full space group G one needs a basis
set which contains basis functions for every point of the star of k. These
will be defined now. We denote the number of star points by s ( k ) . With
p ( k ) as the order of the small point group of k, we have s ( k ) = p / p ( k ) .
From the full point group P of equivalent directions we select s ( k ) elements
a1 = E , 122, . . . , a S ( k ) each generating a different star point k l , k 2 , ..., k s ( k ) .
Thus, we can write alk k = k l , a 2 k = k 2 , . . . , a , ( k ) k = k , ( k ) . The
choice of elements a j , j = 1,2, . . . , s ( k ) is not unique because, together with
a j , each element generated by multiplication with an element ai of the
small point group P k j of k j forms another element which transforms kl into
k j . The choice of these elements made at the outset will be maintained
throughout.
The basis functions of an irreducible representation of the small space
group Gkj will be denoted by ( p k j 1 , ( p k j 2 , .. . ,( p k j d . They can be generated
from the basis functions p k n ( x ) , n = 1,2, . . . ,d, of the irreducible repre
sentation D k of the small space group G k in the same way in which, for
non-symmetric k , the basis functions C p k j ( x ) were obtained from the ( p k ( X )
using relation (A. 174) above. Accordingly, we set

One can easily show that the functions ( p k 3 n ( X ) ,n = 1 , 2 , . . . ,d , defined this


way, actually form a basis of an irreducible representation of the small space
group G k 3 if the ( p k n ( X ) , n = 1 , 2 , . . . ,d do so for the small space group G k
as we assume.
We will prove now that the d x s(k)-dimensional space R k d spanned
by the basis functions ( p k j n ( x ) , ? 2 = 1,2,. . . , d , g = 1,2,. . . ,s(k), of equa-
tion (A.187) forms an irreducible representation space of the full space
group G . To this end, we decompose the full point group P into the
small point groups P k 3 of the various star points k 3 . we denote the el-
ements of P k l by a l l , ~ ~ 1 2. ., ,.a l p ( k ) . Using the latter elements, the el-
ements of P k 3 may be written in the form a 3 1 = ff3 . a 1 1 . a3-',a32 =
a3 . a 1 2 . a3 , . . . , a j p ( k ) = a3 . ' Y l p ( k ) . a3 . The products a3 . alz . a;'
-1 -1

cover the whole point group of equivalent directions P if g and i take all
A.9. Irreducible representations of space groups 705

allowed values, i.e. j = 1 , 2 , . . . , s(k), and i = 1 , 2 , . . . ,p(k). If, in the prod-


ucts a j . ali . a;' the left factor aj is replaced by a y , and j' and i are
varied within their respective definition regions at a fixed value of j , then
one likewise obtains all elements of P . Each element a of P may, therefore,
-
be written in the form a = ajl . a l i . c y j ', where j can be arbitrarily chosen
and j ' and i are determined by a and j uniquely. An analogous statement
holds for the elements of the space group. They allow for a representation
of the form

(A.188)

with a fixed value of j which can be arbitrarily chosen and values of j' and i
depending on g . The lattice translation ta;iR in the second angular bracket
of expression (A.188) guarantees that the lattice translation described by the
whole product (A.188) becomes t R , as it has to be. If g in the form (A.188)
is applied to one of the basis functions qkjn(x), then one obtains by means
of (A.180) and (A.187), the relation

This means that the resulting functions are again in the space R k d spanned
by the basis functions (pk,n(x),n = 1 , 2 , .. . , d , j = 1 , 2 , .. . , s(k). Moreover,
the general definition (A.46) of the operators 9 ensures that the group mul-
tiplication relations 91.92 = g1 . 92 for any two elements g 1 , g 2 of the space
group G hold also in RM. This means that the space R k d forms a repre-
sentation space. It is irreducible, if the partial basis set for the particular
star points k used at the outset, gives rise to an irreducible projective rep-
resentation of the small space groups Gk, as we have supposed. Thus, our
assertion above is proved.
The question of whether one obtains all irreducible representations of a
space group for a symmetrical wavevector if one proceeds in the way de-
scribed above, is answered in the affirmative: By taking all stars {k} of the
first B Z and all irreducible projective representations Dkv of the small point
groups of the respective k-points with the crystallographic factor system, all
irreducible representations D{k}vof the space group follow.
For the character X { k } y ( g ) of an element g = tR++) . a in the space group
representation D{klv, equation (A.189) yields
706 Appendix A . Group theory for applications in semiconductor physics

Here ali is that element of the small point group of k = kl associated with
the element a through the relation ali = a;' . a . a j . Furthermore, j is
defined by the equation ak = kj, and the operators aj are the special star-
generating elements introduced above for the point group P of equivalent
directions, and s(k) is the number of different points of the star {k}. With

(A.191)

as the character of the projective representation of the small point group of


k one obtains from (A.190)

Because of the commutativity of operators under the trace operation, X k v ( a l i )


may be replaced by X k v ( a ) .
In section A.10, we will write down explicitly all irreducible projective
representations of the small point group, with the crystallographic factor
system, for the space groups of the diamond and zincblende structures.

A.9.5 Spinor representations of space groups


The representations of space groups derived above are ordinary or vector
representations in the Hilbert space of scalar functions. They are character-
ized by a star {k} of wavevectors and associated with a particular projective
representation of the small point group of k with the crystallographic factor
system. If one considers space group representations in spinor space instead
of ordinary function space, the only aspect which changes is the meaning of
the projective representations of the small point groups - in the presence of
spin their factor system is given by the product of the crystallographic factor
system and the factor system of the spinor representation discussed in sec-
tion A.6. Projective representations with such a factor system can again be
constructed by means of the double group method. Here one needs projective
rather than ordinary representations of the small double point group with
the crystallographic factor system. The extra projective representations of
the small double point group are the spinor projective representations of
the original small point group. Instead of the double point group method
one may also directly use the general method of section A.7 for constructing
projective representations. This method covers any factor system, includ-
ing also the product factor system of the crystallographic and spinor factor
systems of interest here. As the factor systems and pertinent projective rep-
resentations of point groups are well-known (see Bir and Pikus, 1974), one
only needs to inspect to the corresponding tables. In section A.10 we give
A.9. Irreducible representations of space groups 707

the projective representations of the small point groups for crystals with
diamond and zincblende structure both with and without spin.

A.9.6 Implications of time reversal symmetry


The implications of time reversal symmetry for the irreducible representa-
tions of space groups have already been explored in general terms in section
A.8. The procedure discussed there, i.e. the Herring criterion in conjunction
with the Frobenius-Schur theorem is, however, very cumbersome - one must
sum over all elements of the space group including the G3 translations in
order to utilize it. Below, we will apply the concrete form (A.189) of the
representations in order to express the general criterion of section A.8 in a
more transparent way to facilitate its use.

S t a r s {k} and {-k}

We consider an irreducible representation v { k } u of the space group G in a


particular space R{k},. From section A.8 we know that the time-reversed
space KR{k}u gives rise to the complex conjugate irreducible representation
VTklu. Owing to the construction of the space group representations, V* CkIv
belongs to the star {-k} of the wavevector -k. There are two possible cases,
case 1 with {k} = {-k} and case 2 with {k} # {-k}.
In case 2 the two irreducible representation spaces R{k}, and KR{k}u
are necessarily linearly independent of each other, i.e. case b) of the general
Herring criterion of section A.8 applies, and an additional degeneracy exists
because of time reversal symmetry. The two representation spaces R{k}y and
KRtklu are joined to the same eigenspace of the Hamiltonian by the time
reversal operation.
In case 1, one cannot immediately tell whether the two representation
spaces R{k},, and KR{h}, are linearly independent or not. In order to deter-
mine this, the Frobenius-Schur theorem must be applied.

Theorem of Frobenius and Schur for space groups

For the application of this theorem, the characters of the elements g 2 of


the space group G have to be calculated for a particular irreducible repre
sentation z){k},. According to relation (A.45), the square of a space group
element g = t ~ + ( .~a )is given by the relation

(A.193)

We apply this result to expression (A.192) for the characters X { q u ( g 2 ) of


representations D{k}p It follows that
708 Appendix A. Group theory for applications in semiconductor physics

Summing over the entire space group yields

(A.195)
The R-sum extends over all lattice points of the periodicity region and the
a-sum over all elements of the point group P of equivalent directions. The
R-sum differs from zero only for those a and j for which

a-lk, = -k, + K, (A.196)


holds with K, being a reciprocal lattice vector. The condition (A.196) may
be interpreted in a simple way. For iixed 3 and varying a , the vector a-lk,
runs through all points of the star of k, and, since k, belongs to the same
star as kl k, through all points of the star of k. The condition (A.196) can,
therefore, be fulfilled if among the various points k of the star {k} there are
such which are equivalent to -k. This means that -k belongs to the star of
k, or that the stars of k and -k are identical. This is case 1)defined above.
If the condition (A.196) cannot be fulfilled, then there is no point in the star
of k equivalent to -k, i.e. the stars of k and -k are different. The latter
is case 2) from above. In this case the two representation spaces R { k } , and
K R { k } yare linearly independent, and are joined by time reversal symmetry
to an eigenspace of the Hamiltonian with the same energy eigenvalue. The
degree of degeneracy is doubled, and the representation in KR{k}, forms
the complex conjugate D;,},of the representation D{+ in R{+. While
the star of D:,+ is {-k}, the projective representation of the small point
group of -k belonging to D:,}, is Z)kv* = D;,. If D,& is real, then the
two energy eigenvalues at k and -k, which are degenerate because of time
reversal symmetry, belong to the same projective representation of the small
point group of k, and if z)ky is complex, these eigenvalues belong to different
projective representations complex conjugate to each other.
In case 1) the evaluation of the R-sum yields the value G3 for elements
a which obey condition (A.196). For all other a this sum is zero. One thus
may restrict the a-sum for a particular value of J to those elements a which
obey relation (A.196). We denote the set of these elements a by "5. With
this, equation (A.195) becomes
A.9. Irreducible representations of space groups 709

Since the ali run through the set Mkl if the a vary over a set Mki, the expres-
sion under the j- sum in (A.197) is independent of j . Moreover x k v ( a f i )
may be replaced by x k , ( ( r 2 ) as has already been mentioned above. One
therefore gets

where Mk is the set of all (r from P such that


-
a 'k=-k+K, (A.199)
K being a reciprocal lattice vector.
In case 2) the R sum yields zero, whence

cx(I,>v(s2)
= 0. (A.200)
9
Using the last two relations, the left hand side of the Frobenius-Schur theo-
rem (A.163) for representations of space groups becomes

1 Ip(k)]-' CaEMk
e-iK'7'(a)Xkp(Lu2) {k} {-k}
- cx{I,}u(g2)
=
N L l
{k} # {-k)
1 0 (A.201)
Here, we have used the relation N = G3 . s(k) .p(k) for the total number
N of space group elements. For {k} # {-k} one obviously has the Herring
case b ) , as anticipated at the outset. For {k} = {-k}, all three Herring
cases are possible, depending on which of the relations

is satisfied. As above, one has K 2 = 1 for vector representations, and K 2 =


-1 for spinor representations.
The set Mk does not form a group in general, for a2 is generally not
+
contained in MI, if a does (since a-2k = k - K a-lK). The whole set M I ,
710 Appendix A. Group theory for applications in semiconductor physics

Table A.4: Characterization of the three Herring cases for space groups.

Herring Characteristics
case

a)
{k} = {-k} Ip(k)]-l e-iK'P(at X k u ( a 2 ) = K 2

{k} f {-k} no further condition

{k} = {-k} [p(k)]-l xaEMke-iK'P(a) X k , ( a 2 ) = - K 2

can, however, be generated from a single element a 0 of Mk by multiplying


it with the elements of the small point group Pk of k; thus MI, = ao. Pk. If
k and -k are equivalent, then a0 may be taken as the identity element, and
one has Mk = 9. For inequivalent k and -k, one can take a0 in form of
the inversion I , provided it belongs t o P . Then one has

Mk = I ' Pk. (A.203)


Using the above results we will now discuss the implications of time reversal
symmetry for the energy bands E,(k) in a more explicit form.
Additional degeneracy of energy bands

The characterization of the three Herring cases for space group representa-
tions is summarized in Table A.4. Unlike the Frobenius-Schur theorem in
its general form (A.136), the sums in Table A.4 are carried out only over the
subset Mk rather than over the whole space group. Let us first consider a
star with {k} = {-k}.
{k} = {-k}.
In cases b) and c) time reversal symmetry joins the irreducible represen-
tations R{k}, and K R { k } , = R{k}fi into one eigenspace of the Hamiltonian.
One therefore has E,(k) = E,(k), meaning that each energy band is at
least 2-fold degenerate because of time reversal symmetry. In case b), the
two bands E,(k) and E p ( k ) belong to different irreducible representations
A.9. Irreducible representations of space groups 711

of the space group, and in case c) to the same one. For case a), p formally
becomes v, and the relation E,(k) = E,(k) is a tautology, devoid of new
information. For stars with {k} # {-k},

time reversal symmetry joins the irreducible representation spaces R { k ) ,


and KRjk},, = R { - k } , into one eigenspace of the Hamiltonian. w e have
E,(k) = E,(-k), which means that the two bands Ey(k) and E,(k) have
equal energies in different parts of the first B Z . The two bands, therefore,
overlap, but need not be doubly degenerate as in the case above.

Example

We demonstrate the application of the Herring criterion using the example


of a non-symmetric point k from the interior of the first BZ. The small point
group of k contains, therefore, only the identity element. Two cases have
to be distinguished: i) the complete point group P of equivalent directions
contains the inversion I , and ii) P does not contain I .
In case i) the star {-k} of -k is identical to the star {k} of k. In order
to use the Herring criterion, the sum over the set M k needs to be calculated.
In the present case Mk consists only of the inversion. The reciprocal lattice
vector K in (A.199) is zero, and the only representation of the small point
group is the identity representation with Xku(0") = 1. The sum upon MI,
therefore yields 1. Without spin this corresponds to the Herring case a),
i.e. there is no additional degeneracy. With spin the Herring case c) is
obtained in which an additional degeneracy exists. It occurs between the
two Bloch spinors p k and K p k = pi both of which belong to the only spinor
representation of the point group Pk = E , but they do this in orthogonal
spaces. The two bands E(k) and E'(k) belonging to pk and K p k z q$ ,
respectively, are therefore identical. There is a 2-fold degeneracy of all bands
E(k) in the presence of spin, if spatial inversion symmetry and time reversal
symmetry exist. If spin-orbit interaction is omitted, the degenerate states
p k and pk may be interpreted as 'spin up' and 'spin down' states.

In case ii), i.e. without inversion, one has {-k} # {k}. The Herring case
b) applies, and the two Bloch functions or spinors ' p k and Kpk = ' p I k belong
to the same energy: we have E(k) = E'(-k) where again E(k) and E'(k)
are, respectively, the energies of the states p k and pk. This means that both
with and without spin, time reversal symmetry causes an additional band
overlap.
712 Appendix A . Group theory for applications in semiconductor physics

A.9.7 Compatibility
From the main text we know that the Bloch functions p k u and the pertinent
energy eigenvalues E,(k) are continuous functions of wavevector k within
the first B Z . Therefore the irreducible representations z ) ( k } , , of the space
group in the eigenspace of the Hamiltonian for a given energy eigenvalue
E,(k) must likewise be continuous functions of k. From this observation
one obtains compatibility relations between the small representations v k p
for a symmetry line k and the small representations v k o u for a symmetry
center ko, which lies on this line.
Here, we will study these relations in greater detail. To this end we as-
sume that there is an energy eigenvalue E,(ko) at the symmetry center ko.
The corresponding eigenfunctions should belong to the irreducible repre-
sentation Vh,,of the small point group P k o of ko. The small point group
P k of a symmetry line on which the symmetry center ko is located forms
a subgroup of P b . A particular irreducible representation Vb,,of P b is
therefore simultaneously also a representation of P k , in general, however,
not an irreducible but a reducible one. Let Z ) b / k p , D b / k p ! , . . . be the ir-
reducible parts into which v k o u decomposes as representation of P k . Be-
cause of continuity of the energy eigenvalues and the representations with
respect to k, E,,(ko) must split along the symmetry line k into bands
E,(k), E,!(k), . . ., and their corresponding eigenfunctions belong to the irre-
ducible representations Vb/k,, D b / k p , . . .. One says that the representation
D b v at the symmetry center ko must be compatible with the representa-
tions V b / k , , v b / k , ! , . . . along the symmetry line k through ko. In partic-
ular cases, the set v k , , / k p , v b / k p j , . . . may reduce to the one representation
v b / k p alone. Then no band splitting occurs, and V b / k , is compatible with
PkOU.
Similar results hold for compatibility between representations at a sym-
metry line and a symmetry plane which is bounded by the symmetry line.
In section A.10 the compatibility relations are given for symmetry centers
and lines of crystals with diamond and zincblende type structure.

A.10 Irreducible representations of small point


groups

A.lO.l Character tables


We will exhibit the characters of the projective irreducible representations of
the small point groups for crystals having zincblende and diamond structure
at symmetry points I?, A, X , A, L , C and K of the first B Z . The point groups
A.10. Irreducible representations of small point groups 713

t o be considered are listed in Table A.5. We start with the structure of lower
symmetry, that of zincblende type crystals.

Zincblende structure

The symmetry group of crystals with zincblende structure is given by the


symmorphic space group F13m ( T j ) . The irreducible representations of the
space group are therefore associated with vector or spinor representations of
the small point groups of wavevectors k. The point k = 0 or

has the highest symmetry, namely that of the full point group T d of equiva-
lent directions. The star of I' consists only of I' itself.
The 24 elements of T d belong to 5 classes according to section 2: E , 6S4,
6S2, 8C3, 60~. As before, each class is characterized by a representative
element, the number in front indicating the number of different elements in
the class. Since there are as many inequivalent irreducible representations as
there are classes, the number of these representations equals 6 in the present
case. To satisfy the Burnside theorem two of them must be 1-dimensional,
one 2-dimensional and two 3-dimensional. The two 1-dimensional represen-
tations are denoted by I'l(A1) and I'2(A2), the 2-dimensional one by I'lz(E),
and the two 3-dimensional ones by I'Is(T2) and r25(Tl). The notations in
parenthesis are the ones which commonly are used to describe the symmetry
of localized one-electron states of of molecules and point perturbations in

Table A.5: Small point groups P and number s of star points for symmetric points
k of the first B Z in zincblende and diamond type crystals. Properties of the time
reversed k-point are also indicated.

k r A X A L K C
714 Appendix A. Group theory for applications in semiconductor physics

Table A.6: Irreducible representations of the small point group T,j of ? ! for
zincblende type crystals. Here and below, the vector representations are given in the
upper part of the table, and the spinor representations in the lower part. The right
column gives basis functions transforming according to the irreducible representa-
tions of the same line. The symbols for the representations given in parenthesis are
the ones commonly used in the context of localized oneelectron states of molecules
and point perturbations in crystals.

crystals. The characters of the 5 representations are shown in Table A.6.


Since the character is a class function it suffices to give it for whole classes
of elements.
In Table A.6 the following notations are used:

3T 1 33 i
I--)
22
= -[Iz
dz - iy t ) + 212 .!)I, I--)
22
= -Ix
fi - iy l), (A.205)

11 1
I--)
22
= ---[J?;
I. + iy 1) + Iz t)], I--)2i 2i = -[-I.J3i - iy t) + Iz l)]. (A.206)

For each representation, Table A.6 also gives a basis of linear combinations
of products of the three Cartesian components x,y, z of the position vector
x, or of the components Jz,Jg,Jz of the angular momentum pseudovector
J. According to their interpretation as basis vectors, these components
transform under the action of an orthogonal operator a-' according to the
A.lO. Irreducible representations of small point groups 715

transposed matrix A-' of equation (A.31), therefore, according to the matrix


A itself. For inversion, the transformation is given by the negative of the
identity matrix if one deals with a vector, and by the identity matrix if
one has a pseudovector. If the coordinate axes align with the cubic crystal
directions, then the elements of T d transform the components of the position
vector as shown in Table A.7. Using this table it may easily be verified
that the basis functions of Table A.6 do, indeed, transform according to the
corresponding representations. We demonstrate this using the representation
I'15 as an example. The identity element E transforms zyz into zyz. This
corresponds to the 3-dimensional identity matrix. The trace of this matrix
is 3, in agreement with the character of E in the representation r 1 5 in Table
A.6. In each representation the character of the identity element equals the
dimension of the representation. As a representative of the class 6S4 we take
the element S42.. According to Table A.7, it transforms xyz in Zzy. This
corresponds to a transformation matrix whose zz-element is -1 and whose
yy- and zz-elements vanish. The trace of the matrix is therefore -1, as
expected. The representative Czz of the class 3cz generates zg2, the trace
of the corresponding matrix being -1. The element C3(111) of 8C3 yields
zxy. The matrix of this transformation has trace 0. Finally, the reflection
uZyof the class 60, gives Gxz. The trace of the corresponding matrix is 1.
In all cases the traces agree with the character values of r15 in Table A.6.
In the double group of T d , the 48 elements are distributed among 8
classes, as was shown in section A.8, where Q forms its own class, and the
classes {C3,C ] } and {C4, C2) of the simple group each give rise to two
classes. The double group therefore has 8 different irreducible represen-
tations. Among them are the 5 ordinary representations, which we have
already come to know in the case of the simple group. Consequently, there
are 3 extra representations. Because of the Burnside theorem, the sum of
the squares of the dimensions of these representations must be 24. From this
one may conclude that there must be two 2-dimensional spinor representa-
tions and one 4-dimensional one. The first two are denoted by r6 and r7,
and the last by r8. The pertinent characters are given in Table A.6. If the
irreducible spinor representation D Iof the full orthogonal group is taken as
I
a representation of T d , it remains irreducible and yields r7. By multiplying
2 ) with
~ one of the 5 ordinary irreducible representations, one obtains:
2

rl D~ = rr , r2x D~ = r6, r12x D~ = r8,


r15x D $ x r6+ rs , r25x D~ = r7+ rs . (A.207)

As far as time reversal is concerned, the star {-k} equals the star {k} in
the case of I", and all 8 representations of the space group belong to the
716 Appendix A. Group theory for applications in semiconductor physics

Table A.7: Transformation of the position vector components x,y, z with respect
to the cubic axes under the symmetry operations of point groups
T d and Oh.

Herring case a). This means that time reversal symmetry does not result in
an additional degeneracy of energy bands at I?.

a ) 0 < C < 1 and all points on symmetrically


The points ( 0 , 0 , 2 ~ < /with
equivalent directions are denoted by A. The star of A has 6 points. The
small point group is C2,. This group is Abelian, i.e. each of the 4 elements E ,
~ , forms its own class. Thus, there are 4 different 1-dimensional
C2, C T ~c,,2
irreducible representations A,, A2, A3, A4 as shown in Table A.8. In the
A.10. Irreducible representations of small point groups 717

Table A.8: Irreducible representations of the small point group Caw of A for
zincblende type crystals. In the lower right corner those group elements are given
which transform the basis functions (shown in the last column) according to the
respective irreducible representation. The components z, y, z refer to the cubic axes.

double group one has, in addition, the class Q. The double group therefore
has 5 different irreducible representations (see Table A.8). The one extra
representation As must be 2-dimensional according to Burnside's theorem.
It coincides with the spinor representation V1 of the full orthogonal group
1
taken as a representation of CzV.Moreover, we have

~ A4 x D L= As.
A1 x V l = A2 x Vr.= A3 x 2 ) = (A.208)
The time reversed star {-k} differs from the star {k} in the case of A (see
Table A.5). Thus, all representations of the space group at A correspond
to the Herring case b). Since all 5 representations of the small point group
are real, time reversal symmetry joins a certain representation at k with the
same representations at -k. At A one therefore has

Eai(k) = EAi(-k), i = 1,2,3,4,5. (A.209)

X
The boundary point (0, 0,27r/a)of the first B Z and all symmetrically equiva-
lent points are denoted by X . The star of X has 3 points, because (0, 0,27r/a)
and (0, 0, -27r/a) are equivalent (in the sense of reciprocal lattice transla-
tions). The small point group of X is D M . It contains the 8 elements
E , C2, S4, ,943,
odl, ad2, U21, U22 which form the 5 classes { E } ,{Cz}, ( 5 4 , S,"},
(U21, U22}, {adl, a d z } . Consequently, there are 5 irreducible representations
among which 4 are 1-dimensional, namely XI, X2, X3, X4, and one, X5, is 2-
dimensional (see Table A.9). In the double group one has the class Q in addi-
tion, and, instead of (S4, Sj}, the two classes (S4,QS;} and (QS4, S;}. The
718 Appendix A. Group theory for applications in semiconductor physics

Table A.9: Irreducible representations of the small point group &d of X for the
zincblende structure

16 elements of the double group are distributed, therefore, among 7 classes.


The two extra representations of the double group must be 2-dimensional
according to Burnside's theorem. These are denoted by x6 and X7. The full
orthogonal group representation V 1 , taken as a representation of the point
5
group D u , coincides with X 7 . In addition, the following relations hold:

The time reversed star {-k} coincides with the star {k} of X . All represen-
tations at X belong to the Herring case a), so that no additional degeneracy
occurs due to time reversal symmetry.

The point (./a)(C, C,6) with 0 < C < 1 and all the points on equivalent di-
rections are denoted by A. The star of A has 4 points. The small point group
is C3v. The 6 elements E , C3, Ci, u v l ,av2, a v 3 are distributed among the 3
classes { E } , {C3,C:}, { u v l ,r V 2 , u v 3 } . There are therefore two 1-dimensional
representations, A1 and A2, and one 2-dimensional representation, A3. In
the double group, each of the three classes of the simple group again forms
a class, and each gives rise to an additional class by multiplying it with &.
Thus there are three spinor representations, two 1-dimensional ones, A4 and
As, and one 2-dimensional one, As. Table A.10 shows all six representations.
The spinor representations are p-equivalent to the vector representations. D 1
z
taken as a representation of C3v coincides with As. One has the relations
A. 10. Irreducible representations of small point groups 719

Table A.lO: Irreducible representations of the small point group C3v of A and L
for the zincblende structure

In the case of A, the time reversed star of -k differs from that of k. Thus,
all representations of the space group at A correspond to the Herring case
b). The real representations A1, A2, A3 and at the point k are joined
by time reversal symmetry with the same representations in the point -k,
and the two complex representations A4, A5 at k with the complex conjugate
representation at -k. From this it follows that at A

L
The point L is the point of A with = 1. Therefore it also has 4 star points,
and the small point group remains also the same, i.e. C3,. Everything said
about the representations at A applies, therefore, to L as well (see Table
A.lO). The only difference between L and A is that for L the star {-k}
does not differ from, but coincides, with the star {k}. The representations
151,L2, LJ and Lg correspond to the Herring case a), i.e. no additional de-
generacy occurs between the corresponding bands. In the case of L4 and L5
one has the Herring case b), so that at L the additional degeneracy
720 Appendix A . Group theory for applications in semiconductor physics

Table A.ll: Irreducible representations of the small point group C, of C and


K for zincblende type crystals.

(A .2 13)
occurs because of time reversal symmetry.

<
The point ( 3 ~ / 2 a ) ( < , < , O0) ,< < 1 and points on equivalent directions
are denoted by C. The point with C = 1 is denoted by K. The number of
star points amounts to 12 in both cases, and the small point group is C,.
Each of the two elements E and ah of C, forms its own class. Thus, there
are two 1-dimensional irreducible representations, C1 and Cz or K1 and Kz
(see Table A.ll). In the double group of C, each class of C, gives rise to yet
another class through multiplication by &. This means that there are also
two 1-dimensional spinor representations, C3 and C4 or K3 and K4 . These
are p-equivalent to the two vector representations. If taken on the point
group C,, Vi becomes reducible and decomposes into the 1-dimensional
5
spinor representations C 3 , C4 or K3, K4, respectively.
The star {-k} differs from the star {k} in the case of C and K , so that
the Herring case b) applies. Thus the following relations hold at C:

The same relations also hold at K .


Diamond structure

The diamond structure follows from the zincblende structure if the two prim-
itive f.c.c sublattices are occupied with identical, as opposed to different,
A.lO. Irreducible representations of small point groups 721

atoms. The operation which transforms the two sublattices into each other
thus forms an additional symmetry element of the diamond structure. One
can take it in the form of a glide reflection or the inversion joined with a frac-
tional translation. We choose the second possibility. The inversion is carried
out at the center of the line connecting the two atoms of a primitive unit cell,
and is accompanied by a translation along the connecting line by the amount
a / 4 . The point group of equivalent directions for diamond structure follows,
therefore, from the corresponding point group for the zincblende structure,
i.e. T d , by forming the direct product T d x ci = oh,
As for the zincblende structure, we consider the irreducible representa-
tions of the space group at the symmetric points I?, A, X , A, L , C and K . The
corresponding small point groups are shown in Table A.5. Adding inversion,
the number of star points either doubles in comparison with the point group
of the zincblende structure, or remains the same while doubling the number
of elements of the small point group. For all symmetry points mentioned
above the stars {k} and {-k} are the same in the diamond structure (see
Table A.5).
Due to the occurrence of fractional translations in the space group of the
diamond structure, not only vector and spinor representations of the small
point groups are required, as in the zincblende case, but also projective
representations with other factor systems. Actually, this applies only to
k-points at the boundary of the first B Z , i.e. to X , L , and K , for only
these k-points can have a crystallographic factor system different from the
identity system. As we will see below, only that of X differs essentially from
the identity system, while that of L is p-equivalent to the latter, and that of
K is equal to it. Those of I',A, A and C are equal to the the identity system
anyway.
The small point groups C3wand CzWoccur in both the diamond and zinc-
blende structures. Since the crystallographic factor systems of the respective
symmetric points A, K , and C are the identity systems in the diamond
structure as well, the same irreducible representations of these groups apply
as in the zincblende case.
On the other hand, there are small point groups in the diamond struc-
ture which can be obtained from small point groups in the zincblende struc-
ture by multiplying with Ci. In fact only one of the remaining small point
groups of the diamond structure cannot be generated in this way, namely
the point group C4w. In all other cases this construction is possible, one
has oh = 0 x Ci,D4h = D u x Ci and D u = D3 x Ci. If, in these cases,
the crystallographic factor system is the identity system also in the diamond
structure, then the irreducible representations found for zincblende case can
also be used to obtain the representations for diamond case. Let us consider
r: The small point group oh of I? in the diamond structure follows from
the corresponding small point group 0 of I' in the zincblende structure by
722 Appendix A . Group theory for applications in semiconductor physics

multiplying with the inversion I , and the crystallographic factor system is


the identity systems also for oh. The irreducible representations of Oh may
be obtained from the irreducible representations of 0 by multiplying with
the two irreducible representations r+or r- of Ci. In this, r+denotes the
identity representation of Ci, and emerges from r+by assigning the in-
version to -1 instead of +l. In this way, each irreducible representation of a
small point group Pk in the zincblende structure gives rise to two irreducible
representations of the corresponding point group Pk x Ci in the diamond
structure. The irreducible representations formed in this way are all there
are, for they satisfy the Burnside theorem. There is only one small point
group, namely the symmetry group D4h of X , whose irreducible represen-
tations cannot be obtained in this way, since in this case the factor system
differs substantially from the identity system.

The five irreducible vector representations of T d generate 10 vector repre-


sentations of oh, and the three spinor representations of T d give rise to 6
spinor representations of oh. The 16 representations are listed in Table A. 12.
Vector representations for which the inversion is associated with +1 are de-
noted by the same symbol as for T d . Vector representations with -1 for the
inversion are marked by a prime on the symbol for the corresponding repre
sentation of T d . The two 3-dimensional representations are an exception, in
their case the symbols of the representations are exchanged in the transition
from T d to o h : r15 with +I for I becomes the representation rL5 of oh,
and r15 with -1 for I becomes the representation r 1 5 of oh, while r 2 5 with
$1 for I becomes ri5 and r25 with -1 for I becomes r 2 5 (see Table A.13).
These somewhat strange relations result from the fact that the notations I'15
and r25 were originally introduced for the 3-dimensional representations of
the cubic group 0 rather than the tetrahedral group T d . In the presence
of spin, the representations of oh with fl for I are denoted by an upper
index ' + I , and those with -1 for I by an upper index ' - I on the symbol for
the corresponding representations of T d . Simultaneously, the designations
r6 and r7 are exchanged in the transition from T d to oh.
The representation V1 of the full orthogonal group becomes I:' on oh.
T
In addition, we have the following relations:

r12x D~ = r$ , r:, x v 1 x r, ,
A.lO. Irreducible representations of small point groups

-1 -
a"
0 (3 0 3 1
3 I 3 3 1
3 I 3 3 0 0 0 0 0 01;;
+ b 3

3 3 I 3 I 3 1
3 -I$ $

0- 3 1 3 0 0 3 3 1 3 0 c

0 0 0 0 0 0

Q 3 I 3 0 I 3 3 3 H 0 k3 -

N N V N N C"(h

Table A.12: Irreducible representations of the small point group oh of I? for dia-
mond type crystals.
724 Appendix A. Group theory for applications in semiconductor physics

Td Oh Basis functions

r15 rh5 for I =1 ~ y , y ~ , ~ l :

r15 for I =i c , ~ , z

rz5 ri5 for I=1 J x , J y , Jz

rz5 for I =I X(VU - 12). - ax), Z ( X O - W )


~ ( Z Z

r, r$ for I =1

r; for 1 =i t 1 4$ ) , I & J)>


r7 for I =1 {I T),I 1))
r; for I =i

vr = r; +r, , r;, x ~1 5 = rg+r,+ ,


r15x

r25x vr = rF+r;, r;, x v+= r;+r,+ . (A .2 15)


As far as time reversal symmetry is concerned, all representations at belong
to the Herring case a), so that no additional degeneracy of energy bands
occurs.

A
The star of A has 6 points, and its small point group is C4,,. The 8 elements
are assigned to classes as follows: { E } , {C4,C43}, {Ci}, {cl,c3}, {ff2,c4}.
Thus there are five ordinary irreducible representations, among them four
1-dimensional ones, A l , A2, A3, A4, and a 2-dimensional one, As (see Table
A.14).
In the double group we find the new class {Q}, and the class {C4, C,"}
of the simple group splits into the two classes {C4, Q . Cl} and {Q . C4, Cz}.
There are thus two spinor representations, A6 and A?, of dimension 2. The
representation Vi coincides with A6 on C4,,. We have the relations

A1 X V L =a: X Dl =As, (A.216)


A2 x Vr = A h x Vr = A ? , A5 x V + = & + A ? . (A .2 17)
The time-reversed star {-k} equals the original star {k} in the case of A.
All representations of the space group in A correspond to the Herring case
a), i.e. no additional degeneracy occurs because of time reversal symmetry.
725
A . 10. Irreducible representations of s m d point groups

Table A.14: Irreducible representations of the small point group CdV of A for
diamond type crystals.

The star of X has 3 points. Its small point group is D4h. According to
section A.3, we have D4h = D2d x ci. For elements of the subgroup D 2 d of
D4h the crystallographic factor system w ( q , a 2 ) has the constant value 1
because D2d forms a subgroup of T d and none of the elements of T d is joined
with fractional translations in the space group of the diamond structure.
For the remaining elements, the factors w ( c r l , c q ) generally differ from 1
and must be calculated separately. For a(1,I),for example, it follows that
exp[i(4~/a).(a/4)]= -1. It turns out that the crystallographic factor system
of D a is not pequivalent to the identity system in the case of X and the
diamond structure. It belongs to a class different from that of the identity
system.
In constructing the vector representations of D4hr one may take advan-
tage of the identity D4h = D4 x Ci proved in section A.3. Then the 16
elements of D4h can be associated with the 5 classes of D4 as well as 5 fur-
ther classes, which are generated by the previously mentioned ones through
multiplication by I . Altogether one has, therefore, 10 classes and thus also
10 vector representations of D4h. These are shown in the upper part of Table
A.15 and are denoted by Xp,XH, ..., Xg and Xp', X;', ...,Xg'. For the space
group representations at X without spin the vector representations of D4h
are not helpful, one needs the projective representations of this point group
with the crystallographic factor system discussed above. These are shown
in the middle part of Table A.15. The characters are given for each element
individually because the character is no longer a class function for the pro-
jective representation under consideration. Remarkably enough, none of the
726 Appendix A. Group theory for applications in semiconductorphysics

Table A.15: Projective irreducible representations of the small point group D4h of
X with three different factor systems correspondingto diamond type crystals. The
vector representations (those with unit factor system) are given in the upper part,
the vector representations with crystallographic factor system in the middle part,
and the spinor representations with crystallograhic factor system in the lower part.

-
X E 2c4
-I
xt" 1 1 1

xz" 1 1 1

x; 1 i 1

x; 1 i 1 I 1 1

xs" 2 0 2 0 0 2

x;' 1 1 1 1 1 i

xz"' 1 1 1 I i I

x;' 1 i 1

xi' 1 i i 1 i 1 i
2 0 0 1 0 2 0 0
XgVI -
x1 2 0 0 0 0 0 0 0 0 0 0

x2 2 0 0 0 0 2 0 1 0
x3 2 0 0 0 0 0 0 0 0 0

x4 2
-
0 0 0 0 2 0 2 0

x5 4 0 0 0 0 0 0 0 0
-
E
-I
- C4I c; cir U21I U23' U22I u24r

four representations Xi,X 2 , X3, X 4 is 1-dimensional. This corresponds to


the general rule for irreducible projective representations with a factor sys-
tem which are not p-equivalent to the identity system, mentioned in section
A.7, whereupon 1-dimensional representations are not possible. The Burn-
side theorem allows for yet two possibilities in this case, either there are
four 2-dimensional irreducible representations or a 4-dimensional one. Table
A.15 shows that in the spin-free case considered here the first possibility
applies. For the projective spinor representations of D4h with the crystallo-
graphic factor system, the second possibility is valid - one has exactly one
4-dimensional representation X 5 , which is shown at the bottom of Table
A.15. The product of two projective irreducible representations of D a with
the same factor system decomposes into irreducible vector representations
Of D4h.

The time reversed star {-k} equals the original star {k} in the case of
X . Time reversal symmetry does not result in additional degeneracy at X,
A.lO. Irreducible representations of small point groups 727

Table A.16: Irreducible representations of the small point group D 3 d for diamond
type crystals.

L E 2c3 3U" 1 zc31 3avI Basis functions

L1 1 1 1 1 1 1 *I;zx f uu

L2 1 1 I 1 1 I JZ

L3 2 i 0 2 i 0 { J z - Jz, Ja, - J r )

L; 1 1 1 i i 1 x i l l i =

L; 1 1 i i i 1 .( - U)(# - 2 x 2 - x)

Lb 2 I 0 2 1 0 t=- 2 , u - 2 )

with (xu.) -
Lq-basis from Tab.A.6

(JzJYJz)

with ( z v z ) -
L g - h i s from Tab.A.6

(JzJYJz)

0 0 II t),Ill}
Lq-basis from Tab.A.6

1 Lg-basis from Tab.A.6

1 0

since all representations of the space group belong to the Herring case a).

A
The star of A has 8 points. The small point group is C B ~and , the crys-
tallographic factor system (A.186) equals the unit system, just as in the
case of the zincblende structure. Thus the same irreducible representations
apply as those shown in Table A.9 for the zincblende structure. Unlike
the zincblende structure, however, the time reversed star {-k} of A equals
the original star {k} of k in the case of the diamond structure. The rep-
resentations A l , A2, h3, correspond to the Herring case a), and the two
representations A4, A5 to the Herring case b). The energy levels A4 and As
are therefore degenerate because of time reversal symmetry. One has

EA,(k) = EA,(k) at A. (A.218)

L
The star of L has 4 points and its small point group is D M = D3 x Ci.
728 Appendix A . Group theory for applications in semiconductor physics

Therein elements occur which, like the inversion, are joined with fractional
translations in the space group of diamond type crystals. Some of these
elements, among them again the inversion, transform L into an equivalent,
but not identical point, whereby the crystallographic factor system differs
from the identity system. It can, however, be shown that this system is p -
equivalent to the identity system. The projective irreducible representations
of D3d to be considered here are, therefore, pequivalent to the ordinary
irreducible representations of this point group. Due t o the relation D3d =
D3 x Ci the latter can be generated from the irreducible representations of
D3 by multiplication with the two representations of Ci. The results are
shown in Table A. 16.
The 6 elements of D3 belong to the 3 classes { E } ,{C3, C,},{ U z l , U22, U23}.
In the double group, each of these classes gives rise to a further class. There
fore, two 1-dimensional representations L1, L2, and a 2-dimensional one, L3,
exist without spin, and also two 1-dimensional representations, Lq, L 5 , and
a 2-dimensional one, L6, with spin. In the group D u = D3 x Ci,one has the
vector representations L1, L2, L3, and the spinor representations Lsf, L,f, Lg
corresponding to +1 for the inversion, as well as the vector representations
L i , Lh, L3, and the spinor representations L 4 , L g , L;, corresponding to -1
for the inversion (see Table A.16). The repTesentation D1/2 of the full orthog-
onal group coincides with L$ on D 3 d . In addition, the following relations
hold:

L3 x V + = L z + L.$+ L.$ , L$ x D zI= L 4 + L , + L;. (A.219)

The time reversed star {-k} equals the original star {k} in the case of L.
The representations L1, La, L.3, L i , Lh, L i as well as L$ and L; belong to
the Herring case a), and the spinor representations L z , L,: and L 4 , L c to
the Herring case b). Thus, time reversal symmetry joins L4f, L z as well as
L.4, L , to representations in the same energy eigenspace. We conclude that
at L,

(A.220)

The small point groups of C and K are both CzV,and the stars both have
12 points. The crystallographic factor systems (A.186) are identical and in
both cases equal the identity system. Therefore, only ordinary and spinor
representations have to be considered. We formulate the results for C. The
point group CzV has already been treated as the small point group of A
A.lO. Irreducible representations of small point groups 729

Table A.17: Irreducible representations of the small point groups CzVof C and K
for diamond type crystals.

av2 Basis functions


u

for the zinrblende structure. According to results obtained earlier, there


are four 1-dimensional irreducible vector representations, C1, Cp, C3, C4, and
one 2-dimensional spinor representation, C5 (see Table A.17). The latter
corresponds to the representation D L of the full rotation group considered
as a representation of the subgroup Cpv.
The timereversed star {-k} coincides with the original star {k}, and
the Herring case a) holds for all representations at C. Thus, no additional
degeneracy of energy bands occurs at C due to time reversal symmetry.
Results identical to those for C follow for K .

A. 10.2 Multiplication tables

Zincblende structure: Point group Td

Table A. 18: I' without spin.


730 Appendix A. Group theory for applications in semiconductor physics

Table A. 19: r with spin.

Table A.20: A without spin. Table A.21: A with spin.

Table A.22: X without spin.

Table A.24: A and L without spin.


A. 10. Irreducible representations of small point groups 731

Table A.27: C and K with and without spin.

A. 10.2 Multiplication tables

Diamond structure: Point group Oh

Table A.28: F without spin.


732 Appendix A . Groilp theory for applications in semiconductor physics

Table A.29: I' with spin.

Table A.30: A without spin Table A.31: A with spin.

Table A.32: X without spin.

x: + x; + xs" x3"+ x; + xgu XX' + x;' + xs"' x3"'+ xquI + xs"'


x; + x; + xgv xi' + Xl' 4-x;t xy + xz"'+ XguI
x,.+x;+xs" x;+x;+xs"
xl"+ x; + xs"

Table A.33: X with spin.

1-1 xl"+ xz"+ x; + x; + 2x;+


x1"'+ x;' + x g + xi'+ 2xs"'
A. 10. Irreducible representations of small point groups 733

Table A.34: A without spin. Table A.35: A with spin.


11 I Ad A5 A6 n

Table A.36: L without spin. Table A . 3 7 L with spin.

Table A.38: C and K without spin. Table A.39: C and K with spin.

1-
734 Appendix A . Group theory for applications in semiconductor physics

A. 10.3 Compatibility relations

Zincblende structure
A. 10. Irreducible representations of small point groups 735

Diamond structure
736 Appendix A . Group theory for applications in semiconductor physics

r; -A; r; ~ h2 r; - c2 L 1 ~

L2 -
r; ~ A; r; - ra ~ c3 L3-
L;

<
~

r;,< A; r;,- A3 r;, c2 La___


Ah c3 Lg -
737

Appendix B

Corrections to the adiabatic


approximat ion

In this Appendix we estimate the corrections terms (2.24) and (2.25) of


section 2.2 to the adiabatic approximation. The difficulty of this estimation
is that the eigenfunctions of the total crystal Hamiltonian $(x, X)$(X)are
not explicitly given; one knows only certain general properties. First of
all, the functions $(x,X),being eigenfunctions of a Hermitian Hamiltonian,
may be assumed to be mutually orthogonal, i.e., for two different functions
$(x, X)and $'(x, X)one has the relation

($'(X)l$(X)) / d3x$'*(x, X)$(x,X) = 0, $' # $. (B.1)

In section 2.2 it is verified that the factors @(X)at @(x, X)d(X)in the total
wavefunctions $(x,X)$(X)obey a Schrodinger equation with a Hermitian
Hamiltonian as well. Thus they also may be assumed to be orthogonal,

(4'14) = 0, 4 # 4. (B.2)
A second property of the total eigenfunction $(x, X)$(X)follows from the
X)+ Vcc(X)
invariance of the crystal potential Vee,ec(~, under a common
translation of the spatial coordinates of all electrons,

and all atomic cores,

x --+ X + t = (XI+ t,Xz+ t , .. . , X J +t ) , 03.4)


thus a displacement of the whole crystal, through the vector t. Then the
eigenfunctions $(x, X)$(X)of the crystal Hamiltonian can simultaneously
738 Appendix B. Corrections to the adiabatic approximation

be chosen as eigenfunctions of the total momentum operator xi +cj


pi Pj.
We will do so and also assume that the crystal as a whole is at rest, so that

The third property we will use involves strong localization of the wavefunc-
tions +(X)at atomic cores. This follows from the large masses of cores - in
order for the kinetic energy operator Tc = &(l/Mj)P! in the Schrodinger
equation (2.25) to compete with the potential energy term, large second
derivatives of the core wavefunction +(X), i.e., strong spatial changes, are
necessary. These can be achieved only through strong localization of these
functions, because the normalization condition requires that d(X) shall de-
cay as X approaches infinity. In the case of a crystal, the core wavefunc-
tions +(X)can have non-zero values only in direct proximity of the lattice
points. In classical terms this corresponds to cores executing small oscilla-
tions around their equilibrium positions.
Fourthly and finally, we use that the electron wavefunction $(x,X),
taken as a function of X,is smooth compared with +(X)which was found
to vary rapidly with X. This may be understood as follows: First, we note
that, because of the small mass of electrons, the argument for strong local-
ization of the core wavefunctions +(X) with respect to core positions X is
not applicable to the variation of the electron wavefunctions +(x, X) with
respect to electron coordinates x. The electron wavefunctions @(x, X), taken
as functions of x,are rather spread out more or less uniformly over the entire
crystal. Second, we form the internal product of relation (B.6) with +(X).
Using the relation
(4IPjld) = 0 (B.6)
which holds because of the oscillatory character of the core motion, we arrive
at

The total number of terms on the left hand side of (B.7) is of the same order
of magnitude as the total number of terms on the right hand side, because
the number of (valence) electrons equals, apart of a factor of the order of
magnitude 1, the number of cores. Moreover, the order of magnitude of the
terms with different i is the same throughout, as is the order of magnitude
of the terms with different j . Therefore, the order of magnitude relations
Appendix B. Corrections to the adiabatic approximation 739

follow for all i, j , indicating that $(x, X)is as smooth with respect to X as
it is with respect to x. From relation (B.8) it follows also that

XI.
P?$(X, X) M Pj2$(X, (B.9)
The four properties of the wavefunctions $(X)and $(x, X)discussed above
will be used to rewrite the two correction terms ($4I4Tc$) and E j ( l / M j ) x
($#lPj$Pj4).The first term has the explicit form

(B.lO)

In rewriting this expression we use the third property concerning the wave
function 4(x) and the fourth concerning the wavefunction $(x, X).Accord-
ingly, the matrix element with respect to the electron states in B.10 depends
only weakly on X, and the factor @*(X)#(X) differs substantially from zero
only in small environments around the equilibrium positions of cores. Thus,
the matrix element can be approximately evaluated at the equilibrium values
of the core coordinates X and be factored out of the integral over X. In this
way one gets

(@+l4Tc$) - (4 14)($ ITcI$ ) .


In a similar way one obtains for the second term
(B.ll)

In relation (B.ll) we replace the kinetic energy operator Tc of cores by the


kinetic energy operator Te of electrons using equation (B.9). This results in
the order of magnitude relation

In a similar way the second term may be rewritten as

Here the product of non-diagonal matrix elements (~,ITe1/21$)(#IT~/214)


has the same order of magnitude as the product ($]T~/21$)(+IT~214) of
the corresponding diagonal elements. According to section 2.2, the kinetic
energy (+IT&) of cores, on statistical average, is smaller than the average
740 Appendix B . Corrections to the adiabatic approximation

kinetic energy ($lTel+) of electrons so that (+ITe)+)sets an upper limit for


(iITe1/21i)(41Tc11/214).
Using (B.13) we arrive at

(B.15)

The two order of magnitude relations (B.14) and (B.15) are the equations
(2.23) and (2.24) used in section 2.2.
741

Appendix C

Occupation number
representation

Within the oneparticle approximation, Slater determinants @vly...vN(x~, x2,


. . . ,XN) of an N-electron system describe stationary many-particle states
composed of oneparticle states vi. In a Slater determinant, a particular
oneparticle state is not ascribed to an individual electron, but all electrons
of the system may be associated with this state with the same probability.
This reflects the principle of indistinguishability of elementary particles in
quantum mechanics. Owing to this principle, a many-electron state cannot
be characterized by statements like electron 1 is in one-particle state v1,
electron 2 is in oneparticle state v2, etc. One can only describe the oc-
cupation of a particular oneparticle state by electrons, without identifying
any individuality of the occupying particles.

Slater determinants in occupation n u m b e r representation

For electrons, a particular one-particle state vi (of definite spin) can either
be occupied with one particle, or not be occupied at all. This observation is
most naturally described by means of an occupation number Nv - if the state
v is occupied, then one has Nu = 1, and if the state is empty, then Nu = 0
holds. For simplicity, we identify the quantum numbers v of one-particle
state with integers i, where i runs from 1 to 00, as there are idnitely many
one-particle states.
The selection {v} = v1,v2,. . . ,UN of one-particle states entering a partic-
ular Slater determinant @iV},determines the entirety of occupation numbers
for all one-particle states, i.e. the sequence N1, N2,. . . ,N,, as follows: the
Ni-value of a state whose quantum number i coincides with one of the quan-
tum numbers vl, v ~ . ,. . , V N ,is 1, and the Ni-value of a state whose quantum
number i differs from all quantum numbers v1, v2,. . . , v ~ is, 0. Conversely,
742 Appendix C. Occupation number representation

a particular sequence N1, N 2 , . . . , N , of occupation numbers Ni of which N


have the value 1and all other the value O, defines a particular Slater deter-
minant of the N-electron system. This means that the many-particle states
of N-electron systems may be represented by vectors I N l , N 2 , . . . ,N,)
whose components are occupation numbers of one-particle states with N
of them having the value 1 and all other the value O, One refers to this
description as the occupation number representation of many-particle states.
Slater determinants @ j u } ( x l x, 2 , . . . ,X N ) describe the same states in coordi-
nate space representation. Each Slater determinant @ { u l ( x l x, 2 , . . . , X N ) is
uniquely assigned to an occupation number vector I N1, N 2 , . . . , N - ) , and,
vice versa, each occupation number vector I N1, N 2 , . . . , N,) corresponds to
a Slater determinant t,b{u}(xl,x 2 , . . . , X N ) :

@{V}(Xl, x2r.. .,X N ) @I N1, N2, . ., N,). (C-1)


As the set of all Slater determinants @ { , } ( x l ,x 2 , . . . ,x ~ forms
a basis in )
the Hilbert space of the N-electron system, the set of all occupation number
vectors I N1, N 2 , . . . , N,) does the same. The space spanned by the occu-
pation number vectors is sometimes referred to as occupation number space
or Fock space.

Creation and annihilation operators

To facilitate use of the occupation number representation, one needs opera-


tors whose effects on the basis functions I N1, N 2 , . . . , N,) of Fock space are
known. To this end one introduces so-called annihilation operators ai and
creation operators a. They are defined by the relations

(N1,N 2 , . . . , 0, . . . , N,), Ni 1,
1

ai I N1, N 2 , . . . , N i , . . . ,N,) =
0, Ni = 0.

0, Ni = 1,
at I N1, N 2 , . . . ,N i , . . . ,N,) =
JN1,N 2 , . . . , 1 . . . ,N , ) , N.j = 0.

According to this definition, ai yields a non-zero state only when Ni = 1,


and the resulting state is one with Ni = 0, while a results only then in a
non-zero state if Ni = 0 , and the resulting state is one with Ni = 1. The
operator ai, therefore, annihilates a particle in state i, and the operator
a: creates a particle in state i. By means of creation operators, one can
generate any stationary many particle state of the N-electron system from
Appendix C. Occupation number representation 743

the so-called vacuum state I O,O, . . . , O ) =I 0) in which all Ni = 0 are zero.


In the sense of the correspondence (C.l), one may therefore write
N
$v(xl, x2,. . * 7 X N ) * rl[ a& I 0). (C.4)
j=l

Initially, the '+'-symbol in a: means nothing more than an upper index. We


will now show that a: is in fact the Hermitian adjoint of the operator ai,
as the notation suggests. To do so, we consider the operator product a t a i .
From the definitions of ai and a t one obtains the relation

a:ai 1 N1, N 2 , . . . , N i , . . . , N,) = Ni 1 N1, N 2 , . . . ,N i , . . . , N , ) . (C.5)

The states 1 N1, Nq, . . . ,N i , . . . , N,) are therefore eigenvectors of a:ai, and
the particle number Ni of state i is the pertinent eigenvalue. One calls a r a i
the particle number operator of the state i. We abbreviate it by I?i, i.e. we
set

Now we multiply equation (C.4) from the left by ( N i , NB,. . . , N L 1. It


follows that

Using the definition of the Hermitian adjoint operator (a+)h.a.of a:, we may
write

For relation (C.6) to be valid, = ai must hold. From this it follows


that a t = i.e. a: is in fact the Hermitian adjoint operator of ai.
The ai and aT do not commute with each other. Simple commutation
relations can be derived in terms of the anticommutator. For two arbitrary
operators A , B the anticommutator [ A ,B]+ is defined as

[ A ,B]+= A B + BA.
With this definition we have
744 Appendix C. Occupation number representation

[a:, 4 + = bij, (C.10)

[Ui, U j ] + = [a:, a;]+ = 0. ((2.11)


These relations can be easily verified by means of the defining equations
((2.2) and (C.3) for, respectively, ai and a'. We leave this to the reader.

Operators in occupation n u m b e r representation

In order to obtain the matrix elements of a quantum mechanical operator Q


in occupation number representation it suffices to express Q in terms of cre-
ation and annihilation operators because the elements of the latter operators
are known by virtue of the defining equations (C.2) and (C.3). We demon-
strate such an expression in terms of creation and annihilation operators for
an operator Q(x1, x 2 , . . . ,XN) which, in coordinate space representation, is
additivily composed of one-particle operators q ( x i ) , so that
N
Q ( x 1 , ~ 2 , .. . , X N ) = c q ( x i ) . (C.12)
i= 1

Some operators of important physical quantities of many-electron systems


are of this form, e.g., the total particle number operator, the total momentum
operator, and the total current density operator. The total energy operator
may be one of them if the particles do not interact mutually (which is not true
for electrons, of course). We will demonstrate that operators Q having the
form (C.8) in coordinate space representation, may be written as operators

in occupation number representation.


To prove this assertion it suffices to demonstrate that the matrix elements
of & ( X I , x 2 , . . . ,XN) between Slater determinants $ $ v l ( ~ xl ,2 , . . . , XN) and
$ ~ y ~ (xx2 ~ . ., XN), are the same as the matrix elements of Q between the
,.,
pertinent occupation number vectors 1 Ni,N;, . . . , N k ) and 1 N 1 , N 2 , . . . ,
Nm). We have to prove, therefore, that the identity

is valid. As far as the left hand side of this equation is concerned, one
easily verifies that the only non-vanishing matrix elements are those between
Slater determinants which differ solely in one column. This means that in
Appendix C. Occupation number representation 745

(v;,vi,. . . , v h ) only one component, say vi, can have a value different from
v1 in ( v l ,v 2 , . . . , v ~ ) .The non-vanishing matrix element reads

(@{vO I Q I @(v}) (v: I q 1~1). (C.15)


For the two occupation number vectors I N i , N i , . . . , N L ) and I N1, N2,
. . . , N , ) , the above described form of the Slater determinants means Nu,=
1, N,,; = 0, and NLl = 0, N , = 1, while for all other vi the occupation
v1
numbers N , and N , must be equal. With these specified values of the NUi
vi
and NL!, the right hand side of equation (C.14) may be expressed as
I

( N i ,N L , . . . ,N L I a$uu I N i , N 2 , . . . , N m ) . (C.16)
The matrix element of u$uU in (C.16) differs from zero only when v = v1
and v = vi hold. Thus, it follows that

This is the same result as in equation (C.15).


If the two vectors 1 N i , N i , . . . , N L ) and I N1, N 2 , . . . , N , ) differ in more
than one occupation number, then the matrix element on the right hand side
of equation (C.14) vanishes, as does the matrix element on the left hand side
between Slater determinants differing in more than one column. With this
equation (C.14) is proven for all states. The operator Q ( x l , x 2 , .. . , X N ) in
coordinate space representation is in fact given by the operator Q of (C.12)
in particle number representation.
One can similarly proceed in the case of operators of the N-electron
system which are additively composed of two-particle operators q ( x i ,x j ) ,
i.e. which may be written as

The corresponding occupation number representation is


747

Bibliography

General literature on quantum mechanics, solid


state physics and semiconductor physics

Ashcroft, N.W., and N.D. Mermin (1976), Solid State Physics, Holt Reinhart
and Winston, New York.
Blakemore, J.S. (1985), Solid State Physics, Cambridge University Press.
Boer, K.W. (1990), Survey of Semiconductor Physics, Van Nostrand Rein-
hold, New York.
Bonch-Bruevich, V.L., and S.G. Kalashnikov (1982), Halbleiterphysilc, Deu-
tscher Verlag der Wissenschaften, Berlin.
Callaway, J. (1976), Quantum Theory of the Solid State, Academic Press,
New York.
Enderlein, R. und A. Schenk (1991), Grundlagen der H a l b l e i t e p h y s i ~Aka-
demie-Verlag, Berlin.
Harrison, W.A. (1980), Electronic Structure and the Properties of Solids:
The Physics of Chemical Bond, Freeman & Co., San Francisco.
Harrison, W.A. (1980), Solid State Theory, Dover.
Haug, A. (1972), Theoretical Solid State Physics, Pergamon Press, Oxford.
Kittel, C. (1963), Quantum Theory of Solids,John Wiley & Sons, New York.
Kittel, C. (1986), Introduction to Solid State Physics, John Wiley & Sons,
New York.
Landau L.S., and E.M. Lifshitz (1958), Quantum Mechanics, Pergamon
Press, Oxford.
Landoldt-Bornstein (1982), Numem'cal Data and Functional Relationships
an Science and Technology, New Series, Volume 17, Semiconductors, Eds.
Madelung O., Schulz, M., Weiss, H., Springer-Verlag, Berlin, Heidelberg,
New York.
Madelung, 0. (1978), Introduction to Solid State Theory, Springer-Verlag,
Heidelberg.
Messiah, A. (1968), Quantum Mechanics, North -Holland Publishing Com-
pany, Amsterdam.
748 Bibliography

Moss, T.S. (Ed.) (1982), Handbook on Semiconductors, Volume 1-4, North-


Holland Publishing Company, Amsterdam.
Rosenberg, H. M. (1978), The Solid State, Clarendon Press, Oxford.
Schiff, L.I. (1968), Quantum Mechanics, McGraw-Hill Book Company, New
York.
Seeger, K.-H. (1989), Semiconductor Physics, Springer-Verlag, Wien, New
York.
Shockley, W. (1950), Electrons and Holes in Semiconductors, D. van Nos-
trand Co., Inc., New York.
Smith, R. A. (1979), Semiconductors, 2nd ed., Cambridge University Press,
London.
Vonsovsky, S.V., and M.I. Katsnelson (1989), Quantum Solid-state Physics,
Springer Series in Solid State Sciences 73, Springer-Verlag, Berlin, Heidel-
berg, New York, Tokyo.
Wilson, A.H. (1953), The Theory of Metals, Cambridge University Press.
Yu, P.Y., and M. Cardona (1995), Fundamentals of Semiconductor Physics,
Springer-Verlag.
Ziman, J.M. (1972), Principles of the Theory of Solids, Cambridge University
Press, Cambridge.

Chapter I

Adler, M., H. Fritzsche, and S.R. Ovshinsky (Eds.) (1985), Physics of Dis-
ordered Materials, Plenum Press, New York, London.
Bonch-Bruevich, V.L., R. Enderlein, B. Esser, R. Keiper, A.G. Mironov, LP.
Zvyagin (1984), Elektronentheorie ungeordneter Halbleiter, Deutscher Verlag
der Wissenschaften, Berlin.
Braun, C.L. (1980), Organic Semiconductors, in Handbook of Semiconduc-
tors 3,North Holland Publ. Co., Amsterdam.
Elliot, S.R. (1983), Physics of Amorphous Materials, Longman, London,
New York.
Epifanov, G.I. (1979), Solid State Physics, Mir Publishers, Moscow.
Kleber, W. (1970), A n Introduction to Crystallography, Verlag Technik,
Berlin.
Madelung, 0. (1964) Physics of 111-V-Compounds, J. Wiley & Sons, New
York, London, Sidney.
Bibljography 749

Meyer, H. (1974), Organic Semiconductors, Dark and Photoconductivity,


Verlag Chemie, Weinheim.
Morigaki, K. (1995), Physics of amorphous semiconductors, World Scientific,
Morin, F.J., and J.P. Maita (1954), Phys. Rev. 96, 28.
Mott, N.F., and E.A. Davis (1979), Electronic Processes in Non-Crystalline
Materials, Clarendon Press, Oxford.
Stuke, J., and W. Brenig (Eds.) (1974), Amorphous and Liquid Semicon-
ductors, Taylor & Francis, London.
Unger, K. (Ed.) (1979), Verbindungs-Halblezter, Akademische Verlagsge-
sellschaft Geest & Portig K.-G., Leipzig.
Vainshtein, B.K. (1994), Modem Crystallography, Springer-Verlag, Berlin,
Heidelberg, New York.

Chapter I1

Abrikosov, A.A., L.P. Gorkov, and I.E. Dzyaloshinski (1963), Methods of


Quantum Field Theory in Statistical Physics, Dover, New York.
Bechstedt, F. (1992), Adv. Sol. State Phys. 32, 161.
Bib, H., and W. Kress (1979), Phonon Dispersion Relations in Insulators,
Springer-Verlag, Berlin.
Bir, G.L., and G.E. Pikus (1974), Symmetry and Strain Induced Effects in
Semiconductors, John Wiley & Sons, New York. London, Sydney.
Bloch, F. (1928), Z. Physik, 52, 555.
Born, M., and K. Huang (1968), Dynamical Theory of Crystal Lattices,
Clarendon Press, Oxford.
Born, M., and I.R. Oppenheimer (1927), Ann. der Phys. 84, 457.
Brillouin, L. (1953), Wave propagation in periodic structures, Dover, New
York.
Callaway, J. (1964), Energy Band Theory, Academic Press, New York, Lon-
don.
Chadi, D.J., J.P. Walter, M.L. Cohen, Y. Petroff, and M. Balkanski (1972))
Phys. Rev. B 5 , 3058.
Chelikowsky J.R. and M.R. Cohen (1974), Phys. Rev. B 10, 5059.
Economou, E.N. (1979) Greens Functions in Quantum Mechanics, Springer-
Verlag, Berlin.
750 Bibliography

Fetter, A.L. and J.D. Walecka (1971) Quantum Theory of Many-Particle


Systems, McGraw-Hill Book Company, New York.
Froyen, S., and W.A. Harrison (1979), Phys. Rev. B 20, 2420.
Hedin, L., and J. Lundqvist (1971), J. Phys. C 4, 2064.
Herman, F., S. Skillman (1963), Atomic Structure Calculations, Prentice
Hall, Englewood Cliffs.
Jones, H. (1960), The Theory of Brillouin Zones and Electronic States in
Crystals, North Holland Publ. Co., Amsterdam.
Kane, E.O. (1966), in Semiconductors and Semimetals 1,page 75, Eds. R.K.
Willardson and A.C. Beer, Academic Press, New York.
Kohn, W., and L.J. Sham (1965), Phys. Rev. 140, 1130.
Kramer, B., K. Maschke, and L.D. Laude (1973), Phys. Rev. B 8, 5781.
Leite, J.R., B.I. Bennett, and F. Herman (1975), Phys. Rev. B 12, 1466.
Luttinger, J.M., and W. Kohn (1955), Phys. Rev. 97, 869.
Martinez, G., M. Schliiter, and M.L. Cohen (1975), Phys. Rev. B 11,651.
Maschke K. (1971), phys. stat. sol. b 47, 511.
Phillips, J.C. (1973), Bonds and Bands in Semiconductors, Academic Press,
New York.
Rossler, U. (1979), Festkorperprobleme/Advances in SoZid State Physics 19,
J. Treusch (Ed.), Friedr. Vieweg 8.z Sohn, Braunschweig, page 77.
Slater, J.C., and G.F. Koster, Phys. Rev. 94, 1498.
Tsidilkovski, J.M. (1982), Band structure of Semiconductors, Pergamon Press,
New York, London.
Majewski, J.A., and P. Vogl (1987), Phys. Rev. B 35, 9666.

Chapter I11
Section 3.4 Shallow levels

Aggarwal, R.L., and A.K. Ramdas (1965), Phys. Rev. 140, A1246.
Kohn, W. (1957), Solid State Physics 5, 257.
Lipari, N.O., and A. Baldereschi (1978), Sol. State Commun. 25, 665.
Pantelides, S.T. (1978), Reviews of Modern Physics, 50, 797.
Schechter, D. (1962), J. Phys. Chem. Solids 23, 237.
Bibliography 751

Section 3.5 Deep Levels

Alves, J.L.A., and J.R. Leite (1986), Phys. Rev. B 34, 7174.
Barraff, G.A., and M. Schluter (1980), Phys. Rev. B 19, 4965.
Beeler, F., O.K. Andersen, M. Scheffler (1985), Phys. Rev. Lett. 55, 1498.
Beeler, F., O.K. Andersen, M. Scheffler (1990), Phys. Rev. B 41, 1603.
Bourgoin, J., and M. Lannoo (1983), Point Defects in Semiconductors 11:
Experimental Aspects, Springer-Verlag, Berlin, Heidelberg, New York.
Caldas, M., A. Fazzio, A. Zunger (1984), Appl. Phys. Lett. 45, 671.
Chadi, D.J. (1992), in Materials Science Forum, 83-87,447.
Dabrowski, J., and M. Scheffler (1992), in Materials Science Forum, 83-87,
735.
Delerue, C., and M. Lannoo (1994), in Materials Science Forum, 143 - 147,
699.
Fazzio, A., M.J. Caldas, A. Zunger (1985), Phys. Rev. B 32, 934.
Fazzio, A., M.J. Caldas, A. Zunger (1984), Phys. Rev. B 30, 3430.
Fleurov, V.N., and K.A. Kikoin (1994), Transition metal impurities in semi-
conductors, World Scientific.
Grimmeis, H.G., and E. Janzen (1986), in Deep Centers in Semiconduc-
tors, S.T. Pantelides (Ed.), Gordon & Breach Science Publishers, New York,
London, Paris.
Haldan, F.D.M. and P. W. Anderson (1976) Phys. Rev. B 13,2553.
Hjalmarson, P.H., P. Vogl, D.J. Wolford, J.D. Dow (1980), Phys. Rev. Lett.
44, 810.
Jaros, M. (1982), Deep levels in semiconductors, Adam Hilger, Bristol.
Lang, V., and R.A. Logan (1977) Phys. Rev. Lett. 39, 635.
Langer, J.M., H. Heinrich (1985), Phys. Rev. Lett. 55, 1414.
Lannoo, M., J. Bourgoin (1981), Point Defects in Semiconductors I: Theo-
retical Aspects, Springer-Verlag, Berlin, Heidelberg, New York.
Lischka, K. (1986), phys. stat. sol. b 133, 17.
Pantelides, S.T. (Ed.) (1986), Deep Centers in Semiconductors, Gordon &
Breach Science Publishers, New York, London, Paris.
Scherz, U., and M. Scheffler (1993), in Semiconductors and Semimetals 38,
page 1, Eds. R K . Willardson and A.C. Beers, Academic Press, New York.
752 Bibliography

Stoneham, A.M. (1975), Theory of Defects in Solids, Clarendon Press, Ox-


ford.
Stoneham, A.M. (Ed.) (1986), Current Issues in Semiconductor Physics,
Adam Hilger, Bristol, Boston.
Tersoff, J., W.A. Harrison (1987), Phys. Rev. Lett. 58, 2367.
Van der Rest, J., and P. Pecheur, (1983), Physica 116B, 121.
Vogl, P. (1981), Festkorperprobleme/Advances in Solid State Physics 21, J.
Treusch (Ed.), Friedr. Vieweg & Sohn, Braunschweig, page 191.
Watkins, G.D., and S.T. Pantelides (Eds.) (1986), Deep Centers in Semi-
conductors, Gordon & Breach Science Publishers, New York, London, Paris,
page 160.
Watkins, G.D. (1994)) Materials Science Forum, 143 - 147, 699.
Zunger, A.,U. Lindfelt (1983), Phys. Rev. B 27, 1191.
Zunger, A. (1986), Solid State Physics, Eds. H. Ehrenreich, D. Turnbull,
39,275.

Section 3.6 Surfaces

Banyai, L., and S.W. Koch (1993), Semiconductor quantum dots, World
Scientific.
Bechstedt, F., and R. Enderlein (1988), Semiconductor Surfaces and Inter-
faces, Akademie-Verlag, Berlin.
Bertoni, C.M., 0. Bisi, C . Calandra, and F. Manghi (1978), Proc. 13th Int.
Conf. Phys. Semicond., Inst. Phys. Conf. Ser. No. 43.
Chiaradia, P., G. Chiarotti, S. Nannarone, und P. Sassaroli (1978), Sol.
State Comm. 26,813.
Chiarotti, G., S. Nannarone, R. Pastore, and P. Chiaradia (1971), Phys.
Rev. B 4, 3398.
Drathen, P., W. Ranke, and K. Jacobi (1978), Surf. Sci. 77 L162.
Giintherodt, H.-J., R. Wiesendanger (Eds.), Scanning Tunneling Microscopy,
Springer-Verlag, Berlin, Heidelberg, New York.
Kurtin, S., T. C. McGill, and C. A. Mead (1969), Phys. Rev. Lett. 22,
1433.
Lannoo, M., and P. Friedel (1991), Atomic and Electronic Structure of Sur-
faces, Springer-Verlag, Berlin, Heidelberg, New York.
Luth, H. (1995), Surfaces and Interfaces of Solids, Springer-Verlag.
Bibliography 75 3

Monch, W. (1995), Semiconductor surfaces and interfaces, Springer-Verlag.


Pandey, K.C. (1982), Phys. Rev. Lett. 49, 223.
Pollmann, J. (1980), Festkorperprobleme/Advances in Solid State Physics
20, J. Treusch (Ed.), Friedr. Vieweg & Sohn, Braunschweig, page 117.
Quate, C.F. (1986), Physics Today, page 1.
Takayanagi, K. (1984), J. Microscopy 136,287.
Talwar, D.N.,
and C.S. Ting (1992), Phys. Rev. B 52, 2660.
Terzibaschian, T., and R. Enderlein (1986), phys. stat. sol. b 133, 443.

Section 3.7 Semiconductor microstructures

Altarelli, M., U. Ekenberg, and A. Fasolino (1985), Phys. Rev. B 32, 5138.
Ando, T., A.B. Fowler, and F. Stern (1982), Rev. Mod. Phys. 54, 437.
Bastard, G., Wave mechanics applied to semiconductor heterostructures, Les
editions de physique.
Burt, M. G. (1992), J. Phys.: Condens. Matter 4, 6651.
Cho, A.Y., and K.Y. Cheng (1981), Appl. Phys. Lett. 38, 360.
Christen, J., and D. Bimberg (1990), Phys. Rev. B 42, 7213.
Dingle, R., W. Wiegmann, and C.H. Henry (1974), Phys. Rev. Lett. 33,
827.
Dohler, H.G. (1972), phys. stat. sol. b 52, 79 und 533.
Esaki, L., and R. Tsu (1970), IBM J. Res. Develop., January 1970, page 61.
Frank, W. (1981), Festkorperprobleme/Advances in Solid State Physics 21,
J. Treusch (Ed.), Friedr. Vieweg & Sohn, Braunschweig, page 225.
Herman, M. A., and H. Sitter (1989), Molecular Beam Epitazy, Springer Ser.
Mat. Sci. Vol. 7, Springer, Berlin.
Ivchenko, E.L., and G. Pikus (1995), Superlattices and Other Heterostruc-
tures: Symmetry and Optical Phenomena, Springer-Verlag, Berlin, Heidel-
berg, New York.
Jaros, M. (1989), Physics and Applications of Semiconductor Microstruc-
tures, Clarendon Press, Oxford.
Lei, X.L., N.J.M. Horing, and H.L. Cui (1991), Phys. Rev. Lett. 66,3277.
Leonard, D., M. Krishnamurthy, C.M. Reaves, S.P. Denbaars, and P.M.
Petroff (1993), Appl. Phys. Lett. 63,3203.
754 Bibliography

Notzel, R., N.N. Ledentsov, L. Daweritz, M. Hohenstein, and K. Ploog


(1991), Phys. Rev. Lett. 67, 3812.
Rucker, H. (1985), P.H.D. Thesis, Humboldt-University Berlin.
Rucker, H. (1986), M. Hanke, F. Bechstedt, and R. Enderlein, Superlattices
and Microstructures 2, 477.
Schubert, E.F. (1994), in Semiconductors and Semimetals 40, page 1, Eds.
R K . Willardson and A.C. Beer, Academic Press, New York.
Sibille, A., J. F. Palmier, H. Wang, and F. Mollot (1990), Phys. Rev. Lett
64, 52.
Sipahi, G.M., R. Enderlein, L.M.R. Scolfaro, and J.R. Leite (1996), Phys.
Rev. B 53, 9930.
Stutzmann, M.(1995)) phys. stat sol. (b) 192,273.
Weissbuch, C., and B. Winter (1991), Quantum Semiconductor Structures,
FusLdamentals and Applications, Academic Press, New York.
Zrenner, A., L.V. Butov, M. Hagn, G. Abstreiter, G. Bom, andG. Weiman
(1994), Phys. Rev. Lett. 72, 3382.
See also collection of review articles on semiconductor superlattices and
quantum wells (1986), IEEE Quantum Electron. QE22.

Section 3.8, 3.9 Electric and magnetic fields

Aspnes, D.E. (1967) Phys. Rev. 153,973.


Boer, K.W., H. J. Hansch, and U. Kiimmel (1959), Z. Phys. 155, 170.
Cardona, M. (1969), Modulation Spectroscopy, Academic Press, New York.
Dexter, R.N. (1956), H.J. Zeiger, and B. Lax, Phys. Rev. 104,637.
Enderlein, R., and R. Keiper (1967), phys. stat. sol. 19,673.
Enderlein, R (1996), phys. stat. sol. (b) 194,257 (1996).
Franz, W. (1958), Z. Naturforsch. 13a, 484.
Keldysh, V.L. (1958), Sov. Phys.-JETP 7, 778.
Pollak, F.H. (1994), in Handbook on Semiconductors, Ed. M. Balkanski,
North-Holland, New York.
Seraphin, B.O., and R.B. Hess (1965), Phys. Rev. Lett. 14, 138.
Tharmalingan, K. (1963), Phys. Rev. 130, 549.
Bibliography 755

Chapter IV

Blakemore, J.S. (1962), Semiconductor Statistics, Pergamon Press, London.


Fistul, V.I. (1969), Heavily Doped Semiconductors, Plenum Press, New York,
London.
Landsberg, P.T. (1978) , Thermodynamics and Statistical Mechanics, Oxford
Univ. Press.
Landsberg, P.T. (1982), in Band theory and transport properties, Handbook
on Semiconductors 1,Ed. T.S. Moss, North Holland PubL Co., Amsterdam.
Shklowskii, B.I., and A.L. Efros (1984), Electronic properties of doped semi-
conductors, Springer-Verlag, Berlin.

Chapter V

Greenaway, D.L. and G. Harbeke (1968), Optical Properties of Semiconduc-


tors, Pergamon Press, New York, London.
Haug, H., and S. W. Koch (1994), Quantum Theory of the Optical and
Electronic Properties of Semiconductors, World Scientsc.
Hummel R. E. (1985), Electronic Properties of Materials, Springer-Verlag,
Berlin, Heidelberg, New York, Tokyo.
Kao, K.C., and H. Hwang (1981), Electrical Transport in Solids, Interna-
tional Series in the Science of the Solid State, Volume 14, Pergamon Press,
New York, London.
Klingshirn, C.F. (1995) , Semiconductor Optics, Springer-Verlag, Berlin, Hei-
delberg, New York.
Peuker, K. (1982), R. Enderlein, A. Schenk, and E. Gutsche, phys. stat. sol.
b 109, 599.
Ftidley, B.K. (1982), Quantum Processes in Semiconductors, Oxford Univer-
sity Press.
Scholl, E. (1987) , Non-equilibrium phase transitions in semiconductors, Sprin-
ger-Verlag, Berlin.
Singh, J. (1993), Physics of Semiconductors and Their Applications, McGraw-
Hill, New York.
Zawadzki, W. (1980), (1982), in Band theory and transport properties, Hand-
book on Semiconductors 1,Ed. T.S. Moss, North Holland PubL Co., Ams-
terdam.
756 Bibliography

Chapters VI and VII

Auth, J., D. Genzow, K.-H. Herrmann (1977), Photoelektrische Erscheinun-


gen, Akademie-Verlag, Berlin.
Boer, K.W. (1992), Survey of Semiconductor Physics, Vol. II, Van Nostrand
Reinhold, New York.
Ebert, G., K. von Klitzing, C. Probst, and K. Ploog (1982), Sol. State
comm. 44, 95.
Moss, T.S. (Ed.) (1982), Handbook on Semiconductors 4, North-Holland
Publ. Comp., Amsterdam.
Queisser, H.J. (1985), The Conquest of the Microchip, R. Piper, GmbH &
Co., Munchen.
Sah, C.T. (1994) Fundamentals of solid state electronics, World Scientific.
Streetman, B.G. (1980), Solid State Electronic Devices, ed. by N. Holonyak,
Prentice Hall, Englewood Cliffs.
Sze, S.M. (1981), Physics of Semiconductor Devices, 2nd ed., John Wiley
and Sons, New York.
Sze, S.M. (Ed.) (1991) Semiconductor devices: Pioneering papers, World
Scientific.
Volz, H. (1986), Elektronik f u r Naturwissenschaftler, Akademie-Verlag, Berlin.

Appendix A

Heine, V. (1960), Group theory in Quantum mechanics, Pergamon Press,


New York.
Koster, G.F. (1957) Space groups and their representation, in Solid State
Physics 5, page 174, Academic Press, New York.
Ljubarski, G. J. (1962), Anwendungen der Gruppentheorie in der Physik,
Deutscher Verlag der Wissenschaften, Berlin.
Streitwolf, H.-W. (1967), Gruppentheorie in der Festkorperphysik, Akademis-
che Verlagsgesellschaft Geest & Portig K.-G., Leipzig.
M. Tinkham (1964), Group theory and quantum mechanics, McGraw-Hill,
New York.
Weyl, H. (1950), Theory of groups and quantum mechanics, Dover Publica-
tions, New York.
Wigner, E.P. (1959), Group theory and its application to quantum mechanics
of atomic spectra, Academic Press, New York.
757

Index
Abelian group 86, 624, 626, 666 anti-bonding orbital 166, 172, 177
ab initio methods 133 anti-resonance state 301
absorption 41, 433 anticommutator 743
acceptor level 474 antisite defect 228
acceptor transition 304, 321 APW method 139
acceptor 271, 276, 277, 279, 490, arsenic 44
507 artificial semiconductor microstruc-
accumulation layer 553, 555, 613 ture 402
actinides 231 As-antisite defect 331
activation enthalpy 239, 240 As-rich surface 388
active region 600 associates 232
additional degeneracy 710 asymmetric dimer model 384
adiabatic approximation 57, 59, 61, atomic core 51, 53, 54, 55, 60, 63,
355, 737 70
adiabatic potential 60 atomic force microscopy (AFM) 373
adiabatic 61, 62, 64 atomic layer 339
Airy-function 436 atomic orbital 135, 140, 141, 285
all electron problem 134 atomic structure 54, 64, 82, 177,
alloy 27, 403 234, 336, 371
a-Sn 29, 210 Au-center 497
amorphous semiconductor 2, 27 Auger recombination 517, 518
amphoteric center 305, 497 augmented plane waves (APW) 139
amplification 592 average value 460
angular momentum basis 192
angular momentum matrix 187 ballistic transport 610
angular momentum operator 681 banddegeneracy 112,116, 117,184
Angular Resolved UPS (ARUPS) band discontinuity 555, 567
374, 379 band edge 539, 549, 550, 551
anisotropic effective mass 273 band index 108
annealing 238 band model 38, 41
annihilation operator 742 band structure 105, 108, 109, 112,
annihilation 506, 527, 529 113, 133, 211
anthracene 28 band-band recombination 517
anti-bonding energy level 291, 292, Bardeens relation 565
3 14 barrier 422
758 INDEX

base 586, 588, 589 Bravais indices 336


basis of crystal 17, 23 Bravais lattice 9, 10, 11, 13, 15, 17,
BenDaniel-Duke boundary condi- 83, 94, 110, 336, 342
tion 417 Bravais type 10
binary compound 3 Brillouin zone 106, 108
binding energy 273, 275, 276, 280, buckling model 375, 377
282 buckliig 383
bipolar generation 511 bulk region 545, 546, 577
bipolar transistor 4, 50, 536, 573, bulk state surface energy band 359
585, 590 bulk state 358, 389
Bloch electron 501, 524 Burgers vector 240
Bloch factor 92,101,129,135,179, buried collector 592
255, 256, 415, 425, 699 Burnside theorem 666, 691
Bloch function 87, 92, 109, 132,
254, 255, 357, 699 capture at deep centers 511
Bloch integral 256 capture center 519
Bloch oscillation 438, 439 capture coefficient 512, 516
Bloch state 120, 437 capture cross section 516
Bloch sum 135, 364 capture mechanisms 517
Bloch theorem 85, 87, 89, 92, 130, capture rate 512, 515, 523
356 capture time 515
blocking direction 585, 591 capture 284, 511
bluegreen laser diode 407 carrier concentration 34, 43, 537
body-centered 10 CD-player 605
Bohr radius 267, 270 cds 222
Boltzmann distribution 42, 461, 480 CdTe 221
Boltzmann equation 500, 501 cell methods 134, 138
bond orbital approximation 171,173 central cell corrections 243, 244,
bonding energy level 291, 292, 314 274, 276, 281
bonding orbital 166, 172, 177 cesium chloride structure 178
Born-Oppenheimer approximation chalcopyrite structure 33
57 channel 620
boron 45, 46, 47 character of representations 663, 664,
Bose distribution function 462 668, 680, 690
bound interface state 390, 391 character table 712
bound state 268 charge carriers 50, 506, 508, 509
bound surface band 361, 374 charge state 251
bound surface state 361, 386, 387 chemical bonding 51, 53, 138, 165,
Bragg reflection line 358 174, 250
Bragg reflection plane 101, 102, 103, chemical potential 458, 461, 479,
104, 105, 358, 410 499, 542, 575
Brattain 536 chemical shift 276, 280
Braun 4 class of conjugate elements 626,665
INDEX 759

class 642, 686 correlation 78, 301, 475


clean surfaces 334, 336 Coulomb interaction 56, 72,165
cleavage plane 387 Coulomb potential 70, 119
cleavage 335 Coulomb repulsion 462, 463, 464,
cluster method 294 465, 466
cluster 294 covalent bonding 174, 175
cohesion energy 175 creation operator 742
collective many-particle excitation critical point 119, 126, 473
78 critical thickness 404, 406
collector 586, 587, 589 crystal class 13, 14, 345
collision term 502 crystal field splitting 222, 324
collisions 501, 524 crystal lattice 6
common-emitter configuration 586, crystal structure 14, 15, 21, 30
590 crystal system 9, 13, 14, 345
compatibility relation 118 crystal without basis 12
compatibility 712, 734 crystal with basis 12
compensation 284, 489, 491, 492 crystal 6
complete compensation 490 cubic crystal system 10
compositional microstructure 402, cubic lattice 10
406 current-voltage characteristic 559,
compositional disorder 226 585, 594, 610
compound semiconductor 29 cyclic subgroup 625
conduction band discontinuity 394, cyclotron motion 455
403 cyclotron resonance 211, 443, 455,
conduction band edge 265 456
conduction band minimum 180
conduction band 39, 41, 42, 134, d-electron 53, 134, 321
138, 267, 271, 472, 473 d-orbital 138
conductivity 1, 36, 505, 525 d-shell230, 232, 307
configuration dependence 78, 301 dangling bond 382
configuration interaction 78, 301, dangling hybrid 286, 377
305, 307 Debye screening length 547, 555
configuration 67, 70,457,460, 462, deep center 281,282,284,299,307,
465 318, 393, 462, 509, 519
conjugate elements 626 Deep Level Transient Spectroscopy
continuity equation 508, 528, 579, (DLTS) 309
601, 602, 603 defect molecule model 285, 287, 312,
continuous group 624, 673 323, 365, 367
copper oxide 4 defect molecule 321
core electron 52, 53, 133, 134, 137 defect structure 32
core state 135 defects 227
corecore interaction 56, 60 degenerate electron gas 615
core 52, 242, 265 Dember effect 595
760 INDEX

density functional theory 79, 136 drain 620


density of states (DOS) 42, 120, drift current 506, 523
123, 124, 126, 127, 128, drift velocity 525
268, 426, 469, 472, 472 drift 506, 508
density of states mass 126 DX-center 233, 235, 333
depletion mode 622 Dyson equation 79
depletion region 545, 546, 577
detailed balance 513 effective density of states 43, 480,
device 490, 499, 535, 548, 567, 573 48 1
diamagnetic 449 effective g-factor 449, 450
diamond structure 15, 17, 18, 19, effective mass anisotropy 274
116, 118, 720 effective mass equation 252, 253,
diamond type semiconductors 148, 258, 259, 266, 272, 418,
198 420, 434, 444, 449, 452,
diamond 3, 16, 29 539
diffusion coefficient 239, 506, 530 effective mass tensor 121, 262
diffusion current 506, 528, 539 effective mass theory 412, 414
diffusion length 529, 530, 581 effective mass 119, 121, 126, 182,
diffusion theory 611 183, 211, 216, 219, 440,
diffusion voltage 541, 542, 561, 576 449, 472
diffusion 239, 502, 508, 527, 528, effusion cell 398
529, 537 Einstein relation 532, 533
Dimer-Adatom-St acking-Fault (DAS) EL2 center 231, 333, 234
model 381, 383 electric field 45, 433, 440, 499, 501,
dimerization 382 506
Diophantin equation 337 electric potential 539, 542
direct gap semiconductor 224, 215 electrical conductivity 34, 35, 41,
direct product 625, 626, 667, 691 49
displacement groups 627 electro-optic frequency 436
distribution function 472, 500 electrochemical potential 533, 552
divacancy 233 electroluminescence display 31
DLTS 310 electron affinity 558
donor level 270, 474 electron concentration 34, 42, 469,
donor transition 303, 321 470, 484, 492
donor-acceptor pair 232 electron diffraction 371
donor 270, 271, 276, 486, 490, 507 Electron Energy Loss Spectroscopy
doping microstructure 402, 408 (EELS) 375
doping 35, 36, 398, 486, 488 electron gas 478
DOS 432,454, 455 electron system 70, 72, 457
double acceptor 464 electron trap 490
double donor 304, 463, 465, 494 electron-core interaction 56, 242
double group 130, 685, 686, 715 electron-electron interaction 70, 72,
double heterostructure 397, 400, 401 178, 245
INDEX 76 1

electron-hole pair 78, 511 Euler angle 130, 631, 632, 679, 683
Electron-Nuclear-DoubleResonance exchange energy 249
(ENDOR) 308 exchange integral 249
Electron-Paramagnetic Resonance exchange potential 77, 133, 247,
(EPR) 308 26 5
electron-phononinteraction 62, 517 exchangecorrelation energy 80
electronegativity 566 exchangecorrelation potential 81,
electronic elementary excitation 65 241, 245, 428
electronic structure of surfaces 363 excitation energy 73
electronic structure 51, 354, 371, excitation level 304
409, 420 excited state 274, 467, 468, 469,
electroreflectance211, 443, 444, 572 493
electrostatic potential 552 excitonic effects 433
elemental semiconductor of group exciton 78
IV 289 extended state 246, 457
elemental semiconductor 4, 28, 29, extended zone scheme 96, 97
31 Extended-State X-ray Absorption
elementary excitation 65, 72 Fine Structure (EXSAFS)
emission coefficient 512 308
emission rate 512 extra representation 686
emission 511 extraction 579
emitter 586, 587, 588, 589 extrinsic semiconductor 35, 43,47,
empirical TB (ETB) method 159, 483, 484
163
empty lattice band structure 214 f-shell230, 232 , 307
empty lattice 110, 113 Fabry-Perot resonator 600
energy band 2,39, 81,98, 108, 116, factor group 625
125, 149 factor system 687, 690, 703
energy gap 39, 41, 104, 108, 134, Faraday 3
219, 255, 403 fcc lattice 11, 13, 113, 118, 131
enhancement mode 622 Fermi distribution 39, 42, 461, 462
enthalpy of formation 236 Fermi energy 40, 58, 461, 478
envelope function equation 258 Fermi gas 479
envelope function 256, 259, 271, Fermi integral 479
278, 414, 435, 443 Fermi level 41, 250, 305, 477, 483,
epitaxial growth 336, 396 484,488,489, 534, 548
EPR 310 Fermi statistics 461
equilibrium distribution function 500 Feynman theorem 64
equivalent representation 660 Ficks law 239, 502
equivalent directions 12, 14 field effect transistor 4, 386, 613,
equilibrium position 54, 63, 64 620
Esaki diode 536, 593 fieId effect 612, 613
Euclidean algorithm 338 fine structure splitting 307
762 INDEX

finite group 624 ground state 70, 73, 274, 279, 468,
first Brillouin zone 98, 106, 107, 469, 491
110, 114, 116, 700 group IV semiconductors 28, 31,
first SL BZ 410, 411 313, 566
flow direction 585 groups 7, 82, 87, 623
Fock space 742 group theory 623
folding 107, 108, 411 GW approximation 134
forbidden zone 39, 108 gyromagnetic ratio 448
forward bias 585
Franz-Keldysh effect 442 Hall constant 49
Franz-Keldysh oscillations 442 Hall effect 43, 44, 46
free carriers 527, 533 Hartree approximation 68, 66, 76
freez-out of carriers 487 Hartree energy 248
Frenkel defect 233 Hartree potential 69, 70, 77, 241,
245, 247, 265, 428
frozen-core approximation 53, 133
Hartree-Fock approximation 68, 76,
full orthogonal group 636
134
full rotation group 636, 673
Hartree-Fock equation 77, 81
heavy hole 189, 197, 204, 210
Ga-rich surface 388
Hellman-Feynman forces 63, 64
GaAs 4, 30, 31, 220, 395, 384, 387, HEMT 407, 557, 574
398, 411, 429, 431 hermiticity 417
gain coefficient 599 Herring cases 710
GaN 30, 31 Herring criterion 695, 707, 711
gap discontinuity 394, 403 heterojunction 388, 390, 394, 396,
GaP 5, 220 535, 536, 549, 550, 552,
gate voltage 621 574
gate 621 heterojunction bipolar transistor (HBPT)
generating elements 637, 638, 639 406
generation center 519 hexagonal lattice 22, 23
generation of free carriers 509 HgTe 209, 221
generation 506 High Electron Mobility Transistor
generators of infinitesimal rotations (HEMT) 402
6 74 Hittorf 3
Ge 2, 29, 36, 218, 444, 456, 488, hole capture 515
492, 497 hole subbands 429
Gibbs free energy 235 hole trap 490
glide-reflection 14,15,632,633,635 hole 46, 47, 48, 428, 429, 489
grain boundaries 240 holohedral point group 8, 14
grand canonical ensemble 458 homomorphism 626
Greens function method 295, 368 Hubbard energy 249,280,305,309,
Greens function 79, 295, 298, 299, 310, 328, 463
369, 370 hybrid orbital 168, 285, 286
INDEX

hydrogen atom 119, 141, 267 international notation 9, 15, 637


hydrogen model 266, 273 interstitial impurity 228, 239
intrinsic concentration 482, 484
ideal surface 340, 341, 343, 354, intrinsic semiconductor 35, 47, 49,
367 482, 483, 484
ideal crystal 5, 50, 225, 226 inverse element 624
ideal semiconductor 470, 472, 482 inversion layer 614, 615, 617, 620
ideal wurtzite structure 25 inversion 7, 200, 204
identity element 623 inverted band edges 408
identity representation 659, 669 ion bombardment 335
impact ionization 440 ion implantation 238, 537
imperfect semiconductor crystal 225 ionic bonding 177, 178
improper rotation 632 ionic conductor 2
impurity atom 35, 50, 226, 230, ionic crystal 179
237, 238, 250, 251, 312, ionic model 324
474, 501 ionization energy 283, 320, 322, 326
indirect gap semiconductor 215,221 ionization level 304, 464
indirect gap 394 iron group 231
infinite group 624 irreducibility 66 1
infinitesimal rotation 674 irreducible crystal slab 348, 354
infrared detector 31 irreducible part of first BZ 111, 117
injection current 580 irreducible representation 86, 95,
injection 510, 571 117, 118, 278, 362, 657,
inner shell electron 51 661, 665, 679, 698, 690,
inorganic semiconductor 28 698, 704, 712
input power 592 iso-energy surface 125, 217
input resistance 592 isocoric impurity atom 242, 244
insulat or-semiconductor junction 535, isomorphic 627
6 12 isomorphism 626
insulator 1, 34, 41 isovalent impurity atom 243, 252,
integrated circuit 622 313
inter-valley coupling 273, 274 iteration 63
interband transition 438
interband tunneling 440 Jahn-Teller distortion 310, 315
interface charge density 571 Jahn-Teller effect 234
interface field 571
interface potential 571 Kane model 199, 200, 209, 264,
interface resonance 390 473, 474
interface state 564, 565, 567, 570, kinetic coefficient 505
614 Kirchhoffs theorem 587
interface 388 KKR method 139
internal photoeffect 595 Kohn-Sham equation 81
internal transition 303 Koopman theorem 73,81,247,302
764 INDEX

Koster-Slater equation 296, 317 long-range potential 243, 244, 252,


Koster-Slater method 295 281
Kronig-Penney problem 422 low index surfaces 344
luminescence diode 50
Lagranges theorem 625 Luttinger parameters 196, 419
Landau level 453 Luttinger-Kohn functions 135, 180,
Landausubband454 260
laser diode 30, 50, 599, 600 Lut tinger-Kohn Hamiltonian 417,
lattice constant 10, 16, 18, 25, 64, 452
403 Luttinger-Kohn model 189,196,198,
lattice matched heterostructure 404 428
lattice misfit 404 Lowdin orbital 144
lattice mismatched heterostructure Lowdin theorem 144
404
lattice mismatch 403 Madelung constant 178, 179
lattice oscillations 64, 65 Madelung energy 178, 179
lattice plane 336, 338 magnetic field 45, 444, 451, 452,
lattice relaxation 234 454
lattice translation 7, 82, 87, 635 magnetic susceptibility 308
lattice 6, 7, 9, 336 magnetoresistance 211
LCAO 135 main groups 230, 238, 311
lead sulfide 4 majority carriers 48, 485, 489, 520,
LEED (Low Energy Electron Diffrac- 522, 537
tion) 372, 373, 386 many-body effects 309, 315
Levinson theorem 268, 270, 301, many-electron effects 284
361, 476 many-electron system 65, 460
lifetime of non-equilibrium electrons many-particle excitations 78
5 14 many-valley semiconductor 215
ligand field theory 324 mass action constant 482
light emitting diode (LED) 30, 31, mass action law 478, 482, 543
604, 605, 606 mass operator 79
light hole 189, 197, 208 mass spectroscopy 308
line defect 240 matching conditions 415
line perturbation 227 material parameter 505
linear approximation 504 mean free flight time 504
Liouville theorem 501 mean free path length 504, 610
liquid semiconductor 2 MESFET 557, 571, 606
LMTO method 139 Metal Insulator Semiconductor FET
local density approximation (LDA) (MISFET) 386, 620, 621,
80, 134 622
local equilibrium 503, 574 Metal Organic Vapor Deposition
localized state 246, 460, 457, 462 (MOCVD) 397
long-range order 5 metal sulfide 4
INDEX 765

metal-semiconductor junction 535, multiply ionizable donor 270, 280,


557, 606 462
metal-semiconductor rectifier 573 multi-vacancy 233
metal 1, 2, 29, 34, 41
metastable interstitial state 332 n-channel MOSFET 622
method of invariants 187, 451 N-electron system 66, 70, 73
microelectronics 4, 29, 593, 622 n-type semiconductor 47, 50, 484,
microstructures 396, 402, 406, 408, 489, 494
409, 414, 420, 431 nanostructures 402
migration 238, 239 naphthaline 28
Miller indices 336
narrow gap semiconductor 224,209,
miniband structure 400
483
miniband 410, 423
nearest neighbor 21
minigap 410, 423, 424
nearly-free-electron approximation
minority carriers 48, 485, 499, 522,
98
537
negative differential drift velocity
mirror plane 9, 20
43 1
mismatch strain 404
negative effective mass 431
mixed crystal 29, 30
negative-U center 310
mobility 34, 35, 525, 526, 527
net capture rate 512
MOCVD 398
modulation doping 556 neutrality condition 478, 491
modulation spectroscopy 211 Newtons law 440
Molecular Beam Epitaxy (MBE) nipi-structures 408
388, 397, 336, 409 noble metal 328
MOMBE (Metal Organic MBE) 399 non-adiabatic 62
momentum relaxation time 524 non-equilibrium carriers 284, 509,
MOSFET 536, 567, 571, 592, 606, 522
620, 621, 622 non-equilibrium electrons 514
Mott relation 563, 565 non-equilibrium holes 522
MTO method 139 non-equilibrium processes 50, 499,
muffin-tin method 134, 138 500, 508
multi-phonon process 517 non-equilibrium state 500, 574
multiband effective mass equation non-ohmic contact 561, 562
259 non-polar surface 387
multiple donor 304 non-radiative recombination 284, 518
multiple heterost ructure 397 non-symmetrical k-vector 95
multiple quantum well 424 non-symmorphic 15, 81, 119
multiplet structure 325 normal divisor 625
multiplication table 623, 643, 644, npn-transistor 585, 586, 587, 590,
729 592
multiply ionizable acceptor 271,280 nuclear reaction 53
multiply ionizable center 494 nucleus 52
766 INDEX

occupation inversion 599, 600 pshell230, 231


occupation number representation p-type semiconductor 47, 50, 489,
458, 741 496
occupation number space 742 palladium group 231
occupation number 458, 741 partial compensation 490
Ohms law 502, 504, 505, 525 Pauli principle 39, 42, 67, 76
ohmic contact 561, 562 PbTe 223, 224
ohmic metal-semiconductor contact Peierls instability 377
6 12 Peierls transition 377
one-electron approximation 355 periodic continuation 113
oneelectron potential 133, 134 periodic potential 94, 100
one-electron Schrodinger equation periodic table 4, 232, 285
241, 355 periodicity condition 93, 656
oneelectron state 74 periodicity region 54, 55, 57, 66,
oneparticle approximation 57, 66, 85, 656
68, 71, 73, 77, 741 persistent photoconductivity 332
one-particle energy 72, 73, 309 perturbation theory 98, 99
one-particle excitation energy 73 phenomenological equation 502, 503
oneparticle excitation 73, 78, 81 phonon assisted tunneling 594
oneparticle Schrodinger equation phonons 65, 501
66, 70, 71, 74, 76, 77 phosphorus 45, 274
one-particle state 67, 70, 71, 76, photo-conduction 49
457, 460, 462 photocell 4
oneparticle wavefunction 66, 67, photocopying 31
71 photocurrent 596
optical absorption spectrum 38 photodetector 50, 407, 574, 598
optical excitation 510 photodiode 598
optical fiber communication 31, 407, photoeffect 595
605 photoemission spectroscopy (PES)
optical transition 432 373, 375
optoelectronics 29 photoluminescene 433
OPW functions 135, 136 photon assisted interband tunnel-
orbital motion 75 ing 442
organic semiconductors 28 photoreflectance 572
orthogonal transformation 7, 12, 130, photothreshold 558
628, 629, 630 photovoltage 596, 597
output power 592 pho t ovolt aic element 598
output resistance 592 pinning of Fermi level 565
oxidation state 250, 251, 324, 331 planar defect 240
plane crystal system 342, 343
p-channel MOSFET 622 plane Bravais lattice 342, 343, 358
p-equivalent 689, 690 plane perturbation 227
p-orbital 138, 143, 150 plane space groups 345, 346, 347
INDEX 767

plasma oscillations 78 quantum well (QW) laser diode 400,


plasmon 78 402, 433, 536, 556
platinium group 231 quantum wire 402
pn-diode 585 quasi Fermi level 582, 595
pn-junction 535, 536, 537, 538, 540, quasi-crystal 8
542, 546, 555, 574, 595 quasi-particle method 134
pnp-transistor 585 quasi-particle 65
point contact rectifier 4 quasi-wavevector 92, 94, 95, 104,
point defect 228 500
point group of equivalent directions QW laser diode 406
14, 345, 637
point group symmetry 343 radiative recombination 284, 517
point group of directions 637 rare earth atom 329
point group 8, 9, 17, 82, 86, 95, rare earths 231, 238
354, 362, 627, 636, 637, real semiconductor crystal 226, 474,
6412, 647 475
reciprocal basis 90, 91
point perturbation complex 232
reciprocal lattice vector 89, 90, 96,
point perturbation 227, 239, 241,
242, 474 101
reciprocal lattice 89, 93, 94, 699
point symmetry operation 8, 636
reciprocal SL 410
point symmetry 7, 9, 83
reciprocal vector space 90
polyacetylene 28
recombination center 519
Poisson equation 507, 508, 546, 554,
recombination coefficient 521
362, 570
recombination current 580, 583
polar surface 387
recombination lifetime 522, 523
precipitate 234
recombination rate 520
primitive crystal slab 339
recombination 50, 517, 575, 578,
primitive lattice vector 6, 8, 17,
579, 583
338, 699
reconstructed surface 348
primitive lattice 6, 7, 9, rectification 535, 557
primitive unit cell 6, 30
rectifier 4, 31, 610
projected bulk band structure 360 reduced zone scheme 96, 97
projective representation 118, 684, reducibility 661
687, 688, 690, 702, 703 reducible representation 661
pseudo-wavefunction 137 Reflection High Energy Electron Diffrac
pseudopotential method 136, 139 tion (RHEED) 373
pseudopotential 137, 138 reflection 14, 632, 633
reflectivity 380
quantized Hall effect 536, 556 reflectance spectroscopy 211
quantum device 574 relaxation 505
quantum dot 402 relaxed surface 348
quantum statistics 479 remote bands 201. 205
768 INDEX

representation matrices 658 self-energy operator 79


representations of groups 656 self-interstitial 228
representations with factor system self-organized growth 402
684, 688 semiconductor device 4, 50, 241
repulsive forces 56 semiconductor heterojunctions 388
resonance state 301 semiconductor heterostructures 443,
reverse bias 585 535
RHEED 398 semiconductor microstructures 402
rigid displacement 7,627,635,654, semiconductor optoelectronics 5
682 semiconductor surfaces 572
rocksalt structure 16, 17, 18, 21, semiconductor 1, 34, 41
22, 178 sensor 574
rotation group 679 shallow acceptor 276
rotation-inversion 7, 14, 640 shallow center 462
rotation-reflection 7, 14, 632, 639 shallow donor 276
rotation 14, 631, 633 shallow level 265, 268, 269, 487
Rutherford Backscattering (RBS) Shockley-Read-Hall recombination
308 518
Rydberg energy 267 Shockley 536
short-range order 5
s-atom 146 short-range potential 244,245, 258,
s-orbital 138, 143, 150 28 1
s-shell 230, 231 Shubnikov-de-Haas effect 2 11
saturation current density 585,610 S i c 31
saturation 487 silicon wafer 490
scanning tunneling microscopy (STM) silver sulfide 3
373, 382 simply ionizable acceptor 465, 466
Schottky approximation 542, 544, simply ionizable donor 465
545, 547, 576 single donor 304
Schottky barrier 559, 563, 564 566, single heterostructure 388, 397
608, 609 SiOz/Si-interface 571
Schottky contact 535, 557, 562, 563, Si02/Si-junction 567
607, 612 Si 2, 4, 29, 36, 37, 38, 188, 189,
Schrodinger equation 57, 59, 62, 212, 213, 375, 490
67, 74, 82, 98 slab method 364
Schonflies notation 637 Slater determinant 76, 78, 301, 305,
Schonflies symbol 640 457, 458, 741
screw-dislocation 28, 240, 241 small point group 116, 118, 131,
screw-rotation 14, 632, 633, 634 362, 702, 703, 712
selection rules 670 smooth potential 252, 258
selenium structure 27, 28 solar cell 50
Se 2, 4, 31, 224 solid state shift 52
self-consistent 74, 133, 134, 428 solid state 5
INDEX 769

solubility 237, 238, 311, 328 239, 242, 250, 265, 269,
sp-bonding elements 230 283, 289, 313, 319
space charge region 539 substitutional RE impurity 330
space charge 507, 508 substrate 397, 620
space group symmetry 343 supercell method 294, 367
space group 14, 15, 17, 81, 82, 85, supercell 294, 367
86, 95, 119, 131, 354, 362, superlattice 397, 400, 421, 424, 536
627, 654, 698, 704, 706 surface antiresonance 361
spin degeneracy 473 surface band structure 365
spin-orbit interaction 74, 75, 129, surface Brillouin zone 35, 358, 359
189, 216, 218, 264, 277, surface energy band 357, 363, 358
448 surface field 572
spin-orbit splitting 193, 194, 219, surface photo-efFects 572
279, 473 surface potential 572
spin-orbit-split band 210 surface reconstruction 350, 352, 353,
spinor function 75 354
spinor representation 130, 131, 682, surface relaxation 349, 350, 386
684, 690, 691, 706, 709, surface resonance 361
714
surface state 361
spinor 77, 128, 129, 130, 131, 682
symmetrical k-vector 95, 116
spin 74, 128, 129, 131, 132
symmetry operation 14, 15, 83
spontaneous emission 602
symmetry point 111, 119, 120
stacking fault 28, 240
symmorphic 15
stacking vector 338
star degeneracy 117
star of wavevector 116, 700, 707, Te 2, 27, 31, 224
710, 713 ternary compound 3
statistical average value 502 tetrahedral semiconductor 238, 315,
statistical correlation 479 320
statistical degeneracy 461 thermal annealing 335
statistical operator 458 thermal donor 233
step dislocation 28, 240, 241 thermionic emission theory 611
stereogram 20, 640, 641, 642 thermodynamic equilibrium 49,457,
stereographic projection 640 458, 499, 501, 503, 530,
sticking coefficient 398 533
stimulated emission 599, 603 thermoelectric properties 32
strained layer 404 thermoreflectance 211
structural defect 227 threshold current density 605
structural perturbation 226 tight binding (TB) method 135, 139
subband 410, 423, 424 tight binding approximation 140,
subgroup 624, 625 148, 170
sublevel 424 tight binding matrix elements 153,
substitutional impurity 227, 238, 159
770 INDEX

tight binding parameters 159, 160, valence band discontinuity 328,391,


169 392, 403
time reversal symmetry 95, 132, valence band edge 266
204, 692, 693, 707 valence band maximum 180
time reversal 693 valence band offset 392
total energy 70, 72, 79, 82, 165, valence band structure 140, 209
173, 174, 175, 303, 355, valence band 39, 42, 46, 184, 188,
490 194, 269, 472, 473
total particle number 458, 469 valence electron 30, 51, 52, 133,
transfer matrix method 370 134, 138, 250
transistor 31 valence shell orbital 142
transition groups 230 valence shell 51
transition metal (TM) 231,232,237, Valley 224, 271
251, 320, 393 van Hove singularity 125, 126, 442
translation group 7, 81, 86, 95, 627, vector potential 444
635, 698 vector representation 674, 684, 686,
translation symmetry 6, 7, 348 709, 714
translation 6, 7, 85, 628, 635 Vegards rule 30, 403, 407
transmission electron microscopy (TEM) von Klitzing 536
373, 382
Wannier functions 296
transparency concentration 600
warping of energy bands 198, 188
tunnel diode 50, 573, 593
Weare-Thorpe Hamiltonian 169
two-particle excitation 78
wide gap semiconductor 483, 523
type I heterojunction 396, 397
Wigner-Seitz cell 7, 97, 98, 105,
type I1 heterojunction 396, 397
114, 138, 700
type I1 misaligned 396 work function 557
type I1 staggered 396 wurtzite structure 16, 17, 18, 23,
type I11 heterojunction 396, 397 24, 31, 178

unified defect-model 566 X-ray diffraction 371, 355


unipolar annihilation 511 zero gap semiconductor 210
unipolar generation 511 zincblende structure 16, 17, 18, 19,
unipolar transistor 50 31, 178, 713
unit cell 6, 11 zincblende type semiconductor 148,
universal tight binding parameters 163, 198
159, 160 ZnSe 491
Ultraviolet Photoemission Spectroscopy ZnS 31
(UPS) 374, 374
11-VI semiconductors 31, 221, 566
vacancy 228, 236, 239, 285, 286, 111-V semiconductors 30, 219, 384,
309, 310 387, 566
vacuum-semiconductor junctions 535 N-VI semiconductors 32, 224
INDEX 771

2 x 1 reconstruction 376
7 x 7 reconstruction 376, 381, 382
IL x m reconstruction 352
sp2-bonding 333, 334
sp3-bonding impurity atom 314
sp3-hybrid orbital 166, 290, 330
sp3-hybrid 165, 334
&doping structures 409, 430
A-line 111, 112
r-point 112, 118
A-line 112
n-bonded chain model 381,378, 377,
379, 380, 381
r-bonded chain 377
k . p-interaction 182
k . p-perturbation 185
k . p-method 179
k . p-perturbation theory 259, 262
k-space 93
k-vector 93, 95, 105
(100) surface 379, 387, 388
(110) surface 384, 387, 388
(111) surface 376
2-dimensional electron gas 536, 552,
620
2D lattice 337
2D quasi-wavevector 356
3d-TM atom 320
4d-atom 320
4f-level 331
4f-orbital329
5d-level 331
5d-TM atom 320

S-ar putea să vă placă și