You are on page 1of 232

UNIVERSIT

E CATHOLIQUE DE LOUVAIN
Faculte des Sciences Appliquees
Unite de Physico-Chimie et de Physique des Materiaux
QUANTUM HALL EFFECT SKYRMIONS:
NUCLEAR MAGNETIC RESONANCE
AND
HEAT CAPACITY EXPERIMENTS
Dissertation presentee en vue de lobtention du grade de
Docteur en Sciences Appliquees
par
Sorin MELINTE
Promoteur:
Prof. Vincent BAYOT
Septembre 2001
UNIVERSIT

E CATHOLIQUE DE LOUVAIN
Faculte des Sciences Appliquees
Unite de Physico-Chimie et de Physique des Materiaux
QUANTUM HALL EFFECT SKYRMIONS:
NUCLEAR MAGNETIC RESONANCE
AND
HEAT CAPACITY EXPERIMENTS
Dissertation presentee en vue de lobtention du grade de
Docteur en Sciences Appliquees
par
Sorin Melinte
Membres du jury:
Jury: Prof. Vincent Bayot, promoteur
Dr. Mladen Horvatic
Prof. Jean-Paul Issi
Prof. Jean-Pierre Michenaud
Prof. Mansour Shayegan
Prof. Piotr Sobieski, president
Septembre 2001
nchinare
mamei mele, Ilinca
n memoria lui Nistor, tatal meu,
surghiunit prin Siberia comunista
lui Carmen si Eusebiu
parintelui meu duhovnicesc, Nicodim
Oration
Il ny a en verite quune seule gehenne pour ceux qui ont mal
vecu: lanonymat, lobscurite et leacement denitif, qui du
Lethe, la rivi`ere dOubli, les conduit au euve sans sourires, puis
les emporte au large, `a locean sans limite et sans fond qui char-
rie tout ce qui na servi `a rien, tout ce qui na rien fait, tout ce
qui est reste sans gloire et inconnu.
Plutarque
V. Bayot
C. Berthier
J.-M. Beuken
J.-P. Colinge
S.M. Girvin
E. Grivei
M. Horvatic
J.-P. Issi
L.-P. Levy
J.-P. Michenaud
I. Pop
M. Shayegan
Abstract
Of all the new physics generated by the highly perfect two-dimensional (2D)
electron systems, the quantum Hall eects (QHEs), integer and fractional,
with their richness and complexity are perhaps the most active and excit-
ing. The fractional QHE is the manifestation of a new state of electron
matter - a peculiar, uniform density incompressible liquid phase, observed
at low temperature (T) and in the presence of a strong magnetic eld (B)
perpendicular to the 2D electron layers. Part of the intellectual fascination
with the QHE phases stems from the challenge of nding new concepts to
describe their properties.
The energy levels available to an electron conned to a 2D layer in a
perpendicular magnetic eld are known as Landau levels. Each Landau
level can accommodate many electrons; dividing the number of electrons
per unit sample area by the degeneracy of a Landau level denes the lling
factor . At a precise value of the magnetic eld ( = 1), the 2D electrons
condense into a ferromagnetic QHE ground state: the electronic system is
an itinerant ferromagnet with a quantized Hall resistivity. The low-energy
electron-spin dynamics of this QHE ferromagnet is extremely unusual be-
cause of the subtle interplay between Coulomb interaction among electrons
and Zeeman coupling of electronic spins to the external B. The elemen-
tary excitations are spin-textured objects, called Skyrmions, which display
complex equilibrium behavior reecting liquid, crystalline and glassy phases.
Qualitatively novel physics arises, moreover, because these topological exci-
tations carry electrical charge and possess an eective spin larger than that
of a single electron.
Thermodynamic measurements on 2D electron systems in the fractional
QHE regime were long considered futile because the eects were thought to
be too small to observe. This thesis aims to revise this pessimistic outlook
by reporting measurements of heat capacity and standard nuclear magnetic
resonance down to very low T. Nuclear magnetic resonance (NMR) and
calorimetric measurements in the mK temperature range are a tour de force
v
of experimental technique and conrm the existence of QHE Skyrmions. In
multiple-quantum-well GaAs/Al
x
Ga
1x
As heterostructures, we found a dra-
matic enhancement of the nuclear spin-lattice relaxation rate around = 1,
that cannot be explained at present without express consideration of Skyrmi-
ons being accommodated in the ferromagnetic QHE ground state. The elec-
tron spin polarization peak (detected in NMR Knight shift measurements)
and the observation of nuclear heat capacity of GaAs quantum wells are con-
sequences of the strong, Skyrmion-mediated hyperne coupling between the
nuclear spin system and 2D electrons in the vicinity of = 1. The B- and
T-dependencies of the nuclear-spin lattice relaxation rate around = 1 seem
to be inuenced by the inhomogeneity of the 2D electron system. The quan-
titative analysis of the nuclear spin-lattice relaxation rate measurements is
more dicult as the interplay between the electron-electron interaction and
disorder is not well understood.
Compelling evidence for QHE Skyrmions comes from tilted magnetic
eld studies. Introducing an in-plane magnetic eld causes a spin phase
transition in the electronic system: Skyrmion-like excitations transform to
single spin-ip excitations above a critical Zeeman energy. We report on
tilted magnetic eld heat capacity and nuclear magnetic resonance measure-
ments that may elucidate the discrepancies in the literature concerning the
range of Zeeman energies over which Skyrmions are the stable excitations
of the = 1 ground state. We found a critical Zeeman energy of 0.04 (in
units of Coulomb energy), consistent with Hartree-Fock calculations which
take into account the nite thickness of the electron layers.
Remarkably, the heat capacity (measured as a function of temperature
near = 1) displays a sharp peak at very low T, suggestive of a Skyrmion
solid-to-liquid phase transition. We performed quasi-adiabatic thermal ex-
periments revealing that the mechanism responsible for the peak in heat
capacity vs temperature is a dramatic enhancement of nuclear spin diusion
across the quantum well-barrier interface.
Finally, QHE excitation gap measurements allow microscopic properties
of Skyrmions (such as the number of encompassed reversed spins) to be
established and explored. Our results reveal that the spin of a thermally
activated Skyrmionanti-Skyrmion pair at = 1 is 9, which corroborates
with both Skyrmion spin measured in conventional single-layer 2D electron
systems and theoretical predictions.
vi
Preface
n sudoarea fet ii tale t i vei manca painea ta, pana te vei ntoarce
n pamantul din care esti luat: caci pamant esti si n pamant te
vei ntoarce.
ntaia carte a lui Moise
To promise at the beginning a nice story is most courtly and fashion-
able. This thesis explores the physics of two-dimensional electron systems
exhibiting the quantum Hall eects (integer and fractional) - an extremely
rich set of phenomena with truly fundamental implications. The quantum
Hall eect (QHE) is a large subject with undened frontiers. This book
introduces the readers to basic experimental and theoretical aspects of the
quantum Hall eect. It cannot and do not survey all important and exciting
topics in this eld. This thesis is intended to cover the subject of QHE
Skyrmions. It does so by providing a close examination of two experiments
of vital importance as, I believe, they prevented Skyrmions to fade away
and furnished ample evidence for their intriguing behavior. It also attempts
to show that the growth of high-quality, multiple two-dimensional electron
systems is one of the main factors that have made possible the rise of novel
low-temperature spectroscopy methods, which have tried, and still try, to
oer deep insights into the QHE physics. The data included in this thesis
has been collected on two GaAs/Al
x
Ga
1x
As multiple-quantum-well het-
erostructures containing two-dimensional electron layers; the stages in the
experimental procedure and data analysis were largely the same for both
samples studied.
Chapter 1 sets the scene. Here I present the samples and basic magneto-
transport experiments. The treatment that I have given to the integral
QHE is necessarily perfunctory; for neither the oscillating density of states
nor the presence of disorder can be detached from the explanation of the
vii
phenomenon. I hope that I have not entirely omitted anything that is es-
sential to its comprehension. Traditionally, the discussion of the fractional
QHE follows after the discussion of the integer QHE. The perspectives on
the fractional QHE are various. Many of them were inspired by the art-
ful Laughlins argument. An overview of the fractional QHE is given from
this point of view. The book is cursory in the treatment of few key topics:
(1) QHE excitation gaps, (2) nite thickness corrections and Landau level
mixing, (3) mixed-spin QHE ground states and spin-reversed quasiparticles,
and (4) QHEs in tilted magnetic elds.
Im asking in Chapter 2 what is meant by QHE ferromagnets and what
QHE Skyrmions are. Although much of the detail of the physics has been
stripped away to get to the essentials, I tried hard to make the answer to
these questions clear and the presentation responsible. This chapter repre-
sents my attempt to understand the seminal ideas and rst steps in building
up the theory of QHE ferromagnets. It does not try to replace the existing
books and review articles on this subject, lled with worthy cogitations and
wisdom. It rather tries to present a thorough account of the theoretical
formalism necessary for understanding the following Experimental Work.
Chapter 3 focuses on heat capacity experiments. Heat capacity mea-
surements of two-dimensional electron systems are among the most chal-
lenging experiments because of the very small electron contribution. The
method of ac calorimetry is commonly applied to systems with small heat
capacities which are dicult to isolate thermally and exhibit small signal-
to-background variations. Thereby, it is perfectly suited to the particular
case of two-dimensional electron systems. This technique has been origi-
nally developed by Dr. Joseph Kung Wang for heat capacity measurements
of two-dimensional electron systems in the integer QHE regime and it was
introduced to me by Prof. Vincent Bayot and Dr. Eusebiu Grivei. The dis-
covery of the giant heat capacity of QHE ferromagnets sprang, ironically,
from our experimental study of the heat capacity of multiple-quantum-well
GaAs/Al
x
Ga
1x
As heterostructures in the integer QHE regime at dilution
refrigerator temperatures. In general, heat capacity is a bulk property with
contribution from all components of a given thermodynamic system. Experi-
mentally, this feature is considered a disadvantage in the study of a sample
containing the lattice, the addenda as well as the two-dimensional electron
layers of interest. However, one may elicit useful knowledge about the elec-
tronic system, even though the nuclear spin system gives the dominant con-
tribution to the measured heat capacity. Prof. Vincent Bayot, Dr. Eusebiu
Grivei, and myself have adapted the relaxation-time heat capacity technique
for specic application when the nuclear spin heat capacity of GaAs quan-
viii
tum wells is orders of magnitude larger than the heat capacity of the two-
dimensional electron system. The observation of the nuclear heat capacity of
GaAs quantum wells around Landau level lling factor = 1 necessarily im-
plies a strong coupling between the nuclear spins and the lattice; we present
here compelling evidence that this coupling is provided by QHE Skyrmi-
ons. Heat capacity experiments are accompanied by thermal measurements
performed on the nuclear spin system of a GaAs/Al
0.3
Ga
0.7
As multiple-
quantum-well heterostructure. To our knowledge, these experiments are the
only ones showing signs of a Skyrmion solid-to-liquid phase transition.
In Chapter 4, standard nuclear magnetic resonance (NMR) spectroscopy
partake in the understanding of QHE Skyrmions. Convincing experimental
evidence for the existence of Skyrmions is locked into optically pumped NMR
experiments which probe the two-dimensional electron spin polarization and
dynamics in the QHE regime. For the standard NMR broadened the horizons
opened by the optically pumped NMR technique, the results presented in
this thesis are weighted against the fundamental work of Prof. Sean Barrett
and collaborators (summarized in Appendix B). The organization and style
I adopted for this chapter are those of a research article: the Introduction,
Theory, Experimental methods, Results and discussion, and the Conclusion
sections follow in logical order. I have tried to present enough material that
any reader will be able to achieve some degree of comprehension on the ap-
plication of standard NMR to the study of two-dimensional electron systems
conned to GaAs quantum wells. At the time of writing the Conclusion,
it is clear that several aspects of NMR experiments described in this thesis
need to be understood better. Future work, as I remark in the Epilogue,
is desirable. The NMR study presented here was performed at Grenoble
High Magnetic Field Laboratory. I thank Prof. Vincent Bayot, Dr. Claude
Berthier, Prof. Laurent-Patrick Levy, and Prof. Mansour Shayegan for initi-
ating this project. Throughout the entire process of ferreting out the NMR
fundamentals, I received helpful insights, invaluable friendship, and encour-
agement from Dr. Mladen Horvatic.
Let me include a few words on the process of writing this book. I would
never have succeeded in attaining the modest level of scientic writing with-
out the extensive and patient help received from Prof. Mansour Shayegan,
Dr. Mladen Horvatic, and Prof. Vincent Bayot. The debts that I owe to
many scholars are enormous. In presenting the inception of QHE Skyr-
mions, I have been largely dependent on the work of Prof. Steven Girvin.
Chapter 3, on heat capacity, is an extensively re-written version of three
Physical Review Letters articles, all of them co-authored with Prof. Vincent
Bayot, Dr. Eusebiu Grivei, and Prof. Mansour Shayegan. Parts of the ma-
ix
terial presented in the Chapter 4, on nuclear magnetic resonance, have been
published in a recent issue of Physical Review B. Writing this book was a
collaborative eort. The friends who have given me helpful criticism and
advice are too numerous to be recorded by name. Perhaps the best advice,
deserving all praise, was the following: The average Ph. D. thesis is nothing
but a transference of bones from one graveyard to another.
x
Contents
nchinare . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii
Oration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xix
List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxii
1 The Quantum Hall Eect 1
1.1 Idealized two-dimensional electron systems . . . . . . . . . . . 1
1.1.1 The free electron . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 Idealized quantum wells . . . . . . . . . . . . . . . . . 3
1.1.3 Landau levels . . . . . . . . . . . . . . . . . . . . . . . 4
1.1.4 The idealized two-dimensional density of states . . . . 5
1.2 Realistic two-dimensional electron systems . . . . . . . . . . . 7
1.2.1 Sample fabrication . . . . . . . . . . . . . . . . . . . . 7
1.2.2 Electrons in GaAs quantum wells . . . . . . . . . . . . 12
1.2.3 The Landau level diagram . . . . . . . . . . . . . . . . 15
1.3 Magnetotransport: Experimental details . . . . . . . . . . . . 16
1.4 Fundamental aspects of the QHE . . . . . . . . . . . . . . . . 20
1.4.1 Measuring the QHE . . . . . . . . . . . . . . . . . . . 20
1.4.2 Integral QHE . . . . . . . . . . . . . . . . . . . . . . . 25
1.4.3 Fractional QHE . . . . . . . . . . . . . . . . . . . . . . 28
1.5 Interlude: Ground state theory at = 1 . . . . . . . . . . . . 33
1.6 Supplemental aspects of the QHE . . . . . . . . . . . . . . . . 35
1.6.1 Finite thickness corrections . . . . . . . . . . . . . . . 35
1.6.2 Landau level mixing . . . . . . . . . . . . . . . . . . . 36
1.6.3 Spin and the fractional QHE . . . . . . . . . . . . . . 37
1.6.4 QHEs in tilted magnetic elds . . . . . . . . . . . . . 39
xi
1.7 Magnetotransport: Results and discussion . . . . . . . . . . . 43
1.7.1 Measurements in tilted magnetic elds . . . . . . . . . 43
1.7.2 QHE excitation gap measurements . . . . . . . . . . . 43
1.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
2 Quantum Hall Eect Skyrmions 51
2.1 General formulation . . . . . . . . . . . . . . . . . . . . . . . 51
2.1.1 Symmetry and ferromagnetism . . . . . . . . . . . . . 52
2.1.2 Dimensionality and spin waves . . . . . . . . . . . . . 54
2.1.3 The abomination of topology . . . . . . . . . . . . . . 57
2.2 QHE ferromagnetism at = 1 . . . . . . . . . . . . . . . . . . 65
2.2.1 Spin-wave excitations . . . . . . . . . . . . . . . . . . 65
2.2.2 Topological excitations (QHE Skyrmions) . . . . . . . 68
2.2.3 Further aspects of QHE Skyrmions . . . . . . . . . . . 72
2.2.4 Electron spin polarization at = 1 . . . . . . . . . . . 76
2.3 QHE ferromagnetism near = 1 . . . . . . . . . . . . . . . . 79
2.3.1 Electron spin polarization near = 1 . . . . . . . . . . 79
2.3.2 Quantum treatment of a Skyrmion . . . . . . . . . . . 81
2.3.3 Skyrmion lattices at zero temperature . . . . . . . . . 83
2.3.4 Skyrmion lattices at nite temperatures . . . . . . . . 86
2.3.5 Skyrmions and the nuclear spin-lattice relaxation rate 86
2.4 Interpretation of the = 1 QHE excitation gap . . . . . . . . 89
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3 Heat Capacity Evidence for Skyrmions 93
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
3.1.1 Small-sample calorimetry . . . . . . . . . . . . . . . . 93
3.1.2 Thermal equilibrium in multiple-QW samples . . . . . 96
3.1.3 Samples for heat capacity experiments . . . . . . . . . 98
3.1.4 Heat capacity: Experimental details . . . . . . . . . . 101
3.2 Calorimetry in the integer QHE regime . . . . . . . . . . . . 106
3.2.1 The steady-state, ac-temperature calorimetric method 107
3.2.2 Results and discussion . . . . . . . . . . . . . . . . . . 109
3.3 The holy nuclear heat capacity . . . . . . . . . . . . . . . . 114
3.3.1 The time-constant calorimetric method . . . . . . . . 114
3.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . 119
3.3.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . 119
3.3.4 Disappearance of Skyrmions . . . . . . . . . . . . . . . 122
3.3.5 Ramications . . . . . . . . . . . . . . . . . . . . . . . 124
3.4 Skyrmion lattices . . . . . . . . . . . . . . . . . . . . . . . . . 129
xii
3.4.1 Heat capacity measurements at very low temperatures 129
3.4.2 The variable nuclear thermal coupling model . . . . . 132
3.4.3 Results of quasi-adiabatic thermal experiments . . . . 134
3.4.4 The crystal of Skyrmions and the nuclear spin diusion136
3.5 The nuclear spin-lattice relaxation rate . . . . . . . . . . . . . 140
3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4 Skyrmions Probed by NMR 145
4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.2 Principles of NMR . . . . . . . . . . . . . . . . . . . . . . . . 149
4.2.1 General aspects of NMR . . . . . . . . . . . . . . . . . 149
4.2.2 Interactions of nuclei with electrons. NMR in metals. 153
4.2.3 NMR in QHE systems . . . . . . . . . . . . . . . . . . 157
4.3 Experimental methods . . . . . . . . . . . . . . . . . . . . . . 161
4.3.1 Experimental setup and the sample . . . . . . . . . . 161
4.3.2 Experimental technique . . . . . . . . . . . . . . . . . 163
4.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . 174
4.4.1 Untilted magnetic eld data . . . . . . . . . . . . . . . 174
4.4.2 Tilted-magnetic eld electron spin polarization . . . . 185
4.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
Epilogue 189
A [Appendix] The Nuclear Spin System in GaAs QWs
and the Schottky Heat Capacity 193
B [Appendix] Optically Pumped NMR Observations 195
Bibliography 199
xiii
xiv
List of Symbols
Here are the main symbols and acronyms used in the present thesis. Consis-
tency with traditional usage has been maintained as far as possible. Vectors
are printed in bold-faced fonts.
2DES two-dimensional electron system
A area
A, / magnetic potential
A, A magnetic vector potential
/
eff
eective hyperne coupling constant
a lattice constant
ac alternating current
B magnetic eld (induction)
B

perpendicular magnetic eld component


B magnetic ux density
C heat capacity
D
n
nuclear spin diusion constant
T(E
F
) density of states at the Fermi level
D 1D: one-, 2D: two-, 3D: three-dimensional
DOS density of states
E energy
E
0
ground state energy
E
QW
0
bottom energy of the ground subband
E
QW
1
bottom energy of the rst excited subband
E

cyclotron energy
E
C
Coulomb energy
E
F
Fermi energy
E
xc
exchange energy
E
Z
Zeeman energy
e absolute value of electrons charge
e

quasiparticle charge
xv
e
z
unit vector for the z-direction
f frequency
FWHM full width at the half maximum
g

electrons eective g-factor in bulk GaAs


g
e
electrons g-factor in vacuum
g ratio of the Zeeman energy to the Coulomb energy
H Hamiltonian operator
h Planck constant
I electric current (rms value of the alternating current i)
I magnitude of the nuclear spin
j electric current density vector
K thermal conductance
K
S
Knight shift
k
B
Boltzmann constant
k wavevector
k
F
Fermi wavevector
l
B
magnetic length
L Lagrangian
LL Landau level
m

electrons eective mass in bulk GaAs


m
e
electrons mass in vacuum
M magnetization vector
m magnetic moment
MBE molecular beam epitaxy
n areal electron density
^ number of particles
NMR nuclear magnetic resonance
NSS nuclear spin system
OPNMR optically pumped nuclear magnetic resonance
P electric power
T electron spin polarization
q wavevector
QHE quantum Hall eect
QW quantum well
R electrical resistance
R
0
zero magnetic eld resistance
R
H
Hall resistance
R
L
longitudinal resistance
xvi
r
s
ratio of the Coulomb energy to the cyclotron energy
r space vector
R
j
jth site on the lattice
R radius
RF radio frequency
rms root-mean-square
s spin quantum number
SdH Shubnikov-de Haas
T (lattice) temperature
T
1
1
nuclear spin-lattice relaxation rate
T
1
2
nuclear spin-spin relaxation rate
T
n
nuclear spin temperature
t time variable
V electric potential dierence
V
d
strength of the disorder potential
V
ee
potential energy for electron-electron interactions
V volume
vs versus
w quantum well width
w
0
rms width of the subband wave function
z
j
complex coordinate of jth particle
magnetic susceptibility

QHE excitation gap at LL lling factor

n
nuclear-spin energy level spacing
static dielectric constant of bulk GaAs

0
disorder broadening
gyromagnetic ratio
thermal conductivity

0
zero magnetic eld mobility

B
Bohr magneton
gradient operator
Landau level lling factor

0
nuclear Larmor pulsation

C
cyclotron angular frequency
magnetic ux

0
magnetic ux quantum
(in-plane) orientation angle
[z] many-body total wave function
xvii

0
(z) ground subband wave function
(z) density prole along z-direction
electrical resistivity

0
zero magnetic eld electrical resistivity

s
electron spin stiness
time constant

0
transport scattering time
electrical conductivity
standard deviation
samples tilt angle
innitesimal quantity
step function
xviii
List of Figures
1.1 QHE magneto-transport measurement geometry . . . . . . . 2
1.2 Single-particle DOS vs energy. Spinless 2D electrons . . . . . 6
1.3 Multiple-QW heterostructure growth sequence (sample M242) 8
1.4 Multiple-QW heterostructure growth sequence (sample M280) 10
1.5 Calculated self-consistent charge distribution within the QW
conning potential (sample M242) . . . . . . . . . . . . . . . 14
1.6 Energy spectrum for 2D electrons conned to GaAs QWs . . 16
1.7 Photographs of the investigated specimen in magneto-transport
studies (sample M280) and Bayotron mixing chamber . . . . 19
1.8 Magnetoresistance of sample M280 . . . . . . . . . . . . . . . 21
1.9 Magnetotransport survey (sample M280) . . . . . . . . . . . . 22
1.10 Magnetotransport survey (sample M242) . . . . . . . . . . . . 23
1.11 Low-magnetic eld magnetotransport overview (sample M280) 24
1.12 Single-particle DOS vs energy. Real 2DES . . . . . . . . . . . 28
1.13 Construction of Laughlin quasiparticles . . . . . . . . . . . . 32
1.14 Numerical simulation for the Landau level crossing . . . . . . 40
1.15 Tilted magnetic eld transport overview (sample M242) . . . 41
1.16 Tilted magnetic eld transport overview (sample M280) . . . 42
1.17 QHE excitation gap analysis (sample M280) . . . . . . . . . . 45
1.18 QHE excitation gap analysis (sample M242) . . . . . . . . . . 46
1.19 Re-entrant behavior of the = 4/3 QHE state (sample M280) 48
2.1 Schematic representation of a spin wave . . . . . . . . . . . . 53
2.2 Schematic representation of a XY vortex . . . . . . . . . . . 58
2.3 Illustration of a Skyrmion spin texture . . . . . . . . . . . . . 63
2.4 Numerically calculated spin-wave dispersion of the QHE fer-
romagnet at Landau level lling factor = 1 . . . . . . . . . 67
2.5 Skyrmions magnetization prole . . . . . . . . . . . . . . . . 73
2.6 Hartree-Fock calculations for Skyrmions at Landau level ll-
ing factor = 1. Bare Coulomb interactions . . . . . . . . . . 74
xix
2.7 Hartree-Fock calculations for Skyrmions at Landau level ll-
ing factor = 1. Realistic Coulomb interactions . . . . . . . . 75
2.8 Predicted temperature dependence of the 2D electron spin
polarization at Landau level lling factor = 1 . . . . . . . . 78
2.9 Skyrmions and the 2D electron spin polarization near = 1 . 80
2.10 Qualitative phase diagram for Skyrmion crystal states . . . . 85
2.11 Predictions on the nuclear spin-lattice relaxation rate around
Landau level lling factor = 1 . . . . . . . . . . . . . . . . . 88
2.12 Measured QHE excitation gap at = 1 (sample M280) . . . . 90
3.1 Thermal diagram for multiple-QW samples . . . . . . . . . . 97
3.2 Photographs of an investigated specimen in heat capacity
studies (sample M280) and tilting stage . . . . . . . . . . . . 100
3.3 Carbon paint thermometers for heat capacity measurements . 103
3.4 Thermal response of multiple-QW samples in ac heat capacity
experiments at low temperatures (samples M242 and M280) . 108
3.5 The measured heat capacity in the integer QHE regime (sam-
ples M242 and M280) . . . . . . . . . . . . . . . . . . . . . . 110
3.6 Line shape of the electronic heat capacity in the integer QHE
regime (sample M242) . . . . . . . . . . . . . . . . . . . . . . 111
3.7 Time dependence of the lattice temperature in relaxation heat
capacity experiments (sample M242) . . . . . . . . . . . . . . 115
3.8 The nuclear heat capacity around = 1 (samples M242 and
M280) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
3.9 Angular dependence of the nuclear heat capacity near = 1.
I (samples M242 and M280) . . . . . . . . . . . . . . . . . . . 120
3.10 Angular dependence of the nuclear heat capacity near = 1.
II (sample M280) . . . . . . . . . . . . . . . . . . . . . . . . . 122
3.11 Disappearance of the nuclear contribution to the measured
heat capacity at high tilt angles (sample M280) . . . . . . . . 124
3.12 The measured heat capacity near = 1 as a function of the
Zeeman energy (sample M280) . . . . . . . . . . . . . . . . . 125
3.13 Heat capacity measurements at intermediate tilt angles (sam-
ple M280) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
3.14 QHE excitation gap and heat capacity measurements at Lan-
dau level lling factor = 4/3 as a function of the Zeeman
energy (sample M280). . . . . . . . . . . . . . . . . . . . . . . 128
3.15 The temperature dependence of the heat capacity near = 1
(sample M242). Untilted magnetic eld data . . . . . . . . . 129
xx
3.16 The temperature dependence of the heat capacity near = 1
(sample M242). Tilted magnetic eld data . . . . . . . . . . . 130
3.17 The temperature dependence of the heat capacity near = 1
(sample M280) . . . . . . . . . . . . . . . . . . . . . . . . . . 131
3.18 The numerically calculated Schottky eect in GaAs QWs
(sample M242) . . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.19 The measured temperature of the lattice during the quasi-
adiabatic thermal experiments (sample M242) . . . . . . . . . 135
3.20 Temperature as a function of time during a heat capacity
experiment in the quasi-adiabatic regime (sample M242) . . . 138
3.21 The determined nuclear spin-lattice relaxation rate in calori-
metric experiments (sample M242) . . . . . . . . . . . . . . . 142
4.1 Pictorial illustration of the NMR setup and the theoretical
NMR line shape . . . . . . . . . . . . . . . . . . . . . . . . . 148
4.2 Schematic diagram of the spin-echo sequence . . . . . . . . . 164
4.3 Typical
71
Ga NMR spectra at low temperatures . . . . . . . . 165
4.4 Low-temperature
71
Ga NMR spectra at = 1 . . . . . . . . . 166
4.5 NMR spectra comparing the line shapes of
71
Ga and
69
Ga . . 168
4.6 NMR determination of the low-temperature
71
Ga nuclear spin-
lattice relaxation rate . . . . . . . . . . . . . . . . . . . . . . 170
4.7 The measured low-temperature
71
Ga nuclear spin-spin relax-
ation rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
4.8 The measured 2D electron spin polarization in the extreme
quantum limit . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.9 Temperature dependence of the electron spin polarization at
Landau level lling factor = 1, 1/2, and 1/3 . . . . . . . . . 176
4.10 Comparison between NMR and OPNMR electron spin polar-
ization results near = 1. . . . . . . . . . . . . . . . . . . . . 177
4.11 Temperature dependence of the electron spin polarization
peak at Landau level lling factor = 1 . . . . . . . . . . . . 178
4.12 Magnetic eld (Landau level lling factor) dependence of the
71
Ga nuclear spin-lattice relaxation rate . . . . . . . . . . . . 181
4.13 Tilt angle dependence of the electron spin polarization . . . . 186
xxi
xxii
List of Tables
1.1 Compilation of various structural and electronic properties of
samples M242 and M280 . . . . . . . . . . . . . . . . . . . . . 11
1.2 Physical parameters of stoichiometric GaAs . . . . . . . . . . 12
1.3 Estimates of relevant energy scales for samples M242 and M280 34
1.4 Zeeman energy contribution to the QHE excitation gap due
to thermally-activated spin reversed quasiparticles . . . . . . 38
1.5 The measured QHE excitation gaps (sample M280) . . . . . . 47
2.1 Calculations of the Skyrmion excitation energy which take
into account both the nite thickness of the electron layer
and the Landau level mixing. . . . . . . . . . . . . . . . . . . 76
3.1 Landmark papers describing calorimetric techniques for small-
sample heat capacity measurements . . . . . . . . . . . . . . 94
3.2 Calorimetric techniques for heat capacity measurements in
the QHE regime . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.3 The low-temperature heat capacity of multiple-QW samples
and details of addenda in heat capacity measurements . . . . 99
3.4 Various properties of materials used for the thermal anchoring
of the sample in heat capacity experiments . . . . . . . . . . 105
3.5 Amplitude of the heat capacity magneto-oscillations in the
integer QHE regime (samples M242 and M280) . . . . . . . . 112
4.1 Details of NMR coils . . . . . . . . . . . . . . . . . . . . . . . 162
4.2 The measured line width of NMR spectra . . . . . . . . . . . 171
A.1 Isotope table for GaAs QWs nuclei . . . . . . . . . . . . . . . 194
B.1 The measured
71
Ga nuclear spin-lattice relaxation rate near
= 1 (OPNMR) . . . . . . . . . . . . . . . . . . . . . . . . . 196
B.2 The measured QWs line width near = 1/3 (OPNMR) . . . 198
xxiii
xxiv
Chapter 1
The Quantum Hall Eect
Si iarasi am vazut sub soare ca izbanda n alergare nu este a
celor iut i si biruint a a celor viteji, si pinea a celor nt elept i,
nici bogat ia a celor priceput i, nici faima pentru cei nvat at i, caci
timpul si ntamplarea ntampina pe tot i.
Ecclesiastul
1.1 Idealized two-dimensional electron systems
It is by experiment that quantum Hall eects (QHEs), integer and fractional,
were discovered
1
. Simple electrical measurements of real semiconductor de-
vices containing two-dimensional electron systems (2DESs) have uncovered
a fascinating low-temperature behavior of the Hall resistance when a strong
magnetic eld is applied perpendicular to the 2DES. For a 2DES device
cut into a standard Hall bridge [Fig. (1.1)] and placed in a magnetic eld
perpendicular to the samples surface (i.e., lying along the z-axis), the Hall
resistance (R
H
) is the ratio of the Hall voltage, measured across the 2DES
device in the y-direction, to the current applied in the x-direction. The
hallmark experimental feature is that the Hall resistance has plateaus over
a nite range of values of the magnetic eld, quantized to R
H
= h/(e
2
).
The quantum number is either an integer or a simple rational fraction with
odd denominator. The QHE is suciently reproducible that it provides an
1
The integer QHE was discovered in 1980 by von Klitzing et al. [108]. The very rst
observation of the fractional QHE was reported two years later by Tsui et al. [102].
1
2 CHAPTER 1. THE QUANTUM HALL EFFECT
invariable reference standard of electrical resistance linked to fundamental
physical constants: h/e
2
= 25812.805 . Advocacy of the quantization of
R
H
is based upon general grounds invoking Landau quantization of the elec-
tron orbits, electron localization, and electron-electron interactions. Why
these still unfolding phenomena are so peculiar and amazing is partly due
to the intimate connection between theoretical interpretations of the ob-
served quantization to various topological and eld theories. We have to
look rst at some preliminary aspects of the QHE before to come to the
part of interest here: the domain of spin-related phenomena in 2DESs.
Figure 1.1: Sample geometries for magneto-transport measurements. (Top)
Crude Hall bridge dened by cleaving the sample to appropriate size. The
current contacts are S and G, while the potential probes are A, B, C, and D,
yielding R
H
= [V
BD
[/I
SG
V
H
/I and R
L
= [V
CD
[/I
SG
V
L
/I. (Bottom)
van der Pauw, square-shaped sample. In a rst experiment, A and B are
the current contacts and C and D are the potential probes yielding the
longitudinal resistance R
L
= [V
CD
[/I
AB
V
L
/I. In a second experiment,
A and C are the current contacts and B and D are the potential probes.
Consequently, the Hall resistance R
H
= [V
BD
[/I
AC
V
H
/I is obtained.
1.1. IDEALIZED TWO-DIMENSIONAL ELECTRON SYSTEMS 3
1.1.1 The free electron
Let no one think that properties of an electron are easy to understand.
The properties of this elementary particle we are most familiar with are
the mass m
e
and the charge e. For small-scale phenomena the electron
behaves quantum mechanically and obeys the principles of relativity; the
quantum mechanics of the free electron was captured by Dirac in his famous
relativistic wave equation. He predicted, for the electron, the existence of
new internal degrees of freedom: the electron has a spin angular momentum
s, dened by the spin quantum number s =
1
2
. It turns out that spin
is connected to statistics and electrons obey Fermi-Dirac statistics. The
electron possess a spin magnetic moment
m
e
= g
e

B
s =
e
s, (1.1)
where g
e
is the free electron Lande g-factor,
B
= e/(2m
e
) is the Bohr
magneton, and
e
is the electron gyromagnetic ratio. To the lowest order in
the ne structure constant = e
2
/(c), quantum electrodynamics predicts
an anomaly for the free electron magnetic moment: g
e
= 2[1 + /(2)].
The g-factor for the free electron characterizes the (Zeeman) coupling of
the electrons spin to a magnetic eld B; the spin Hamiltonian of the free
electron is H
s
= m
e
B.
1.1.2 Idealized quantum wells
We know how to solve few, simple quantum mechanics problems for such an
elementary particle. In the following, we shall restrict our area of inquiry
to the particular case when the electron motion has a two-dimensional (2D)
character. In order to understand this point fully, lets consider the motion
of an electron conned to a one-dimensional (1D) innite square quantum
well (QW). The potential energy is given by
V
b
(z) =
_
for [z[ > w/2,
0 for [z[ < w/2,
where w is the QW width. The eigenenergies of the 1D Schrodinger equation
_


2
2m
e
d
2
dz
2
+V
b
(z)
_
(z) = E(z), (1.2)
describing the quantum mechanical bound motion along the z-axis, are
quoted in all textbooks on quantum mechanics
E
r
QW
=

2

2
(r + 1)
2
2m
e
w
2
, (1.3)
4 CHAPTER 1. THE QUANTUM HALL EFFECT
where r = 0, 1, . . . is a non-negative integer. The discreteness of the energy
spectrum is the key reason for the 2D behavior of electrons conned to QWs.
The normalized even and odd wave functions for [z[ < w/2 are

r
(z) =
_
_
2
w
_
1/2
cos
_
(r + 1)
z
w

, r = 0, 2, 4 . . .
_
2
w
_
1/2
sin
_
(r + 2)
z
w

, r = 1, 3, 5 . . .
It is worth noting that there is at least one bound state, irrespective of
the height of the conning barrier, i.e., innite or nite. The quantum
mechanical motion in the xy-plane is described by the Schrodinger equation


2
2m
e
_
d
2
dx
2
+
d
2
dy
2
_
(x, y) = E(x, y). (1.4)
The energy spectrum is continuous and the wave functions are the well-
known plane waves. In polar coordinates, the complete energy levels are
E
r,k

= E
r
QW
+E
k

QW
=

2

2
2m
e
w
2
(r + 1)
2
+

2
2m
e
k
2
|
, (1.5)
where E
k

QW
is the electrons kinetic energy in the xy-plane and k
|
is the
magnitude of the in-plane wavevector. Obviously, one may associate an
electric subband, which represents the kinetic energy arising from the in-
plane motion of the carrier, with each of the QW bound states. In short, an
electron in the ground electric subband (r = 0) of an innitely deep QW has
an energy E
0
QW
=
2

2
/(2m
e
w
2
) and it is described by the wave function

0
(z) = (2/w)
1/2
cos (z/w). What is meant here by 2D electrons is that
electrons have quantized energy levels in one direction (z-direction), but are
free to move in two dimensions (xy-plane).
1.1.3 Landau levels
Lets focus now on how the energy spectrum of an electron splits into Landau
levels (LLs) in the presence of a magnetic eld. Consider the Hamiltonian
for the free electron in a magnetic eld
H
orbital
=
1
2m
e
(p +eA)
2
, (1.6)
where p is the canonical momentum and A is the magnetic vector potential.
The vector potential can be chosen to lie in the plane perpendicular to the
magnetic eld so that the contributions to the Hamiltonian from motion
1.1. IDEALIZED TWO-DIMENSIONAL ELECTRON SYSTEMS 5
along and perpendicular to the magnetic eld separate. Since the motion
along the magnetic eld direction is unchanged, only the 2D motion of the
electron, in the plane perpendicular to the magnetic eld, remains of inter-
est
2
. In order to x these preliminary ideas in a specic way, you have
to imagine that the magnetic eld is applied along the z-axis (B = Be
z
)
and the motion in the xy-plane is the 2D motion of the electron in the
ground electric subband due to the 1D connement in the z-direction. The
xy-plane is taken to be the complex plane with z being a complex coordi-
nate related to the position vector (x, y) via z = (x +iy)/l
B
. The magnetic
length l
B
= [/(eB

)]
1/2
is the natural length scale in the problem
3
. The
eigenenergies of H
orbital
[Eq. (1.6)] are
E
N
=
c
(N + 1/2);
c
= eB

/m
e
. (1.7)
Here
c
is the cyclotron frequency, N = 0, 1, ... is the LL index, and the
energy zero coincides with the bottom of the ground electric subband. The
energy spectrum is discrete, the manifold of states with energy E
N
[Eq. (1.7)]
constitutes the Nth LL, and E

=
c
is the cyclotron energy. The eigen-
functions are of an especially simple form in the lowest (N = 0) LL

N=0
m
(z) =
z
m
(22
m
m!)
1/2
exp
_

[z[
2
4
_
, (1.8)
where m is the angular momentum index. The magnetic eld couples to
the electrons orbital motion and quantizes its kinetic energy into massively
degenerate LLs. The energy eigenvalues [Eq. (1.7)] do not depend explicitly
on the angular momentum index. Including the spin degeneracy, one obtains
the same degeneracy per unit area for all LLs: 2eB

/h.
1.1.4 The idealized two-dimensional density of states
The preceding discussion has been restricted to a single electron. We con-
sider next the density of states (DOS) of a quantum collection of ^ inde-
pendent 2D electrons of areal density n. First, it makes sense to examine
the DOS at zero magnetic eld. Since we are ultimately interested in the
description of 2DESs conned to QWs, we remark that the DOS could be
decomposed into contributions which arise from dierent electric subbands.
2
Note that the Lagrangian of the problem writes as /
2D
= (m
e
/2)v

ev

, where
v is the velocity of the particle and = x, y.
3
The magnetic length is, in fact, the radius of the cyclotron orbit. We have explic-
itly indicated that it is the perpendicular component of the magnetic which controls the
electrons orbital motion.
6 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.2: The at DOS (per unit area) at zero magnetic eld (panel a)
collapses into a series of massively degenerate Landau levels when a perpen-
dicular magnetic eld B is applied to the 2DES (panel b). A disorder-free,
spin-degenerate situation is illustrated.
It is not dicult to obtain the DOS per unit area associated with one electric
subband [panel (a) to Fig. (1.2)]
T
0
=
m
e

2
, (1.9)
where we have taken into account the spin degeneracy. Thus, each electric
subband contributes the same constant quantity to the total DOS
T
0
2D
(E) = T
0

r
(E E
r
QW
), (1.10)
where (x) is the step function. At zero magnetic eld, DOS exhibits jumps
of nite amplitude T
0
whenever the energy passes through the edge of an
electric subband [panel (a) to Fig. (1.2)]. In the ground state, i.e., at zero
1.2. REALISTIC TWO-DIMENSIONAL ELECTRON SYSTEMS 7
temperature (T = 0), the electrons occupy all states with energy E less
then a limiting value E
F
(the Fermi energy). If n is known and all electrons
can be accommodated into the ground electric subband, using the simple
formula E
F
= n/T
0
, the Fermi energy could be readily obtained.
Consider now the DOS of a 2DES conned to a QW when a magnetic
eld is applied perpendicular to the plane of the 2D electron layer. Assuming
that only the ground electric subband is populated, the DOS per unit area
(including the spin degeneracy) is
T
2D
(E) =
2eB

N
(E E
N
), (1.11)
whit (x) denoting the Dirac delta function. The fact that DOS have -like
singularities centered at the LLs energy eigenvalues [panel (b) to Fig. (1.2)]
witnesses for the remarkable properties of 2DESs placed in a nite, homoge-
nous perpendicular magnetic eld.
1.2 Realistic two-dimensional electron systems
1.2.1 Sample fabrication
Modern QHE experiments are performed almost exclusively on 2DESs con-
ned to GaAs/Al
x
Ga
1x
As heterostructures. The samples investigated in
this thesis were constructed by means of molecular beam epitaxy (MBE).
Initiated in the early 1970s, MBE is presently an extremely versatile tech-
nology for the fabrication of multilayer structures based on lattice-matched
semiconductors such as GaAs and Al
x
Ga
1x
As. In this technique, the crys-
talline structure is obtained via reactions between thermal molecular beams
of elemental species and a substrate surface maintained at a high tempera-
ture in ultrahigh vacuum. Since it is inherently a slow growth process and
electrically active impurities are incorporated to the growing crystal with
separate beams, extreme dimensional control over both compositional vari-
ations and doping prole can be achieved. It is known that MBE produces
atomically smooth interfaces and the conduction and valence bands exhibit
abrupt steps at a common interface between GaAs and Al
x
Ga
1x
As. The
conduction band of GaAs sits lower in energy and the steps are 300 meV
for an Al mole fraction of x = 0.3. In the physical formation of 2DESs
described below only the conduction band prole will be involved.
8 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.3: Design for the multiple-QW sample M242 is shown schematically.
1.2. REALISTIC TWO-DIMENSIONAL ELECTRON SYSTEMS 9
In the conventional (simplest) approach for the realization of a 2DES (see, for
example, Fig. (1.11) in Ref. [86]), the essential part of the MBE grown struc-
ture consists of an undoped GaAs layer, followed by an undoped Al
x
Ga
1x
As
spacer layer and then a Si-doped Al
x
Ga
1x
As region. To maintain a con-
stant Fermi level throughout the structure, excess electrons, donated by the
remote Si impurities, nd their way to the conduction band of GaAs. This
charge transfer creates strong internal electric elds that cause signicant
bending of the conduction band in the vicinity of the GaAs/Al
x
Ga
1x
As
interface, which, in turn, acts as a connement potential for electrons. A
2DES results as the carrier motion on the direction perpendicular to the
interface is quantized. After donating their electrons, the impurity atoms
are left positively charged, the net charge of the sample thus being zero
4
.
The single-QW arrangement could be viewed from the substrate upward
as a combination of an inverted interface (GaAs on top of Al
x
Ga
1x
As) and a
normal interface (Al
x
Ga
1x
As on top of GaAs). For modulation-doped QW
heterostructures, dopants are placed on both sides of the QW [Fig. (1.3)].
Single QW samples and conventional heterointerfaces provide a unique realm
for standard magneto-transport measurements. For thermodynamic studies
on 2DESs
5
, however, one has to seek for a method that will enhance the
measurable signal. One straightforward way to solve this problem is to
construct multiple-QW heterostructures [90].
The two samples used in this thesis were grown by MBE on semi-
insulating, (100)-oriented GaAs substrates at Princeton University. It can
be easily seen in Figs. (1.3) and (1.4), that samples M242 and M280 are sim-
ilarly structured. They are composed of one hundred GaAs QWs bounded
on each side by Al
x
Ga
1x
As barriers which are -doped with Si near their
centers. A complicated sequence of layers was grown on the substrate preced-
ing the multiple-QW system and following it. The rationale behind choosing
such a ponderous design is to fashion 2DESs with very high mobilities [74].
For the purposes of this thesis, I mention here some important character-
istics of the heterostructure design. From the top downward, the structure is
composed of a GaAs cap layer followed by a Al
x
Ga
1x
As layer. This region
consists of two -doped Si layers: the majority of the impurities are placed
4
Although the parent donors are spatially separated from the 2DES by the spacer
layer, they unavoidable contribute to the electron scattering. The remote ionized-impurity
scattering is one of the mechanisms responsible for the nite low-T mobility of the 2DES.
This (zero-magnetic eld) low-T mobility is a major physical variable which assesses the
quality of the sample.
5
Determination of the DOS via specic heat measurements is an example.
10 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.4: Design for the multiple-QW sample M280 is shown schematically.
1.2. REALISTIC TWO-DIMENSIONAL ELECTRON SYSTEMS 11
in a plane far from the 2DES and close to the surface. The -layer situated
closer to the QW provides electrons for the 2DES, while the heavy-doped
-layer mainly passivates the surface states at the GaAs-air interface. The
QWs have been slightly asymmetrically -doped by placing the donors near
the center of the barriers. Asymmetric doping has been used to compensate
for the migration of Si along the growth direction [65]. The amount of Si
dopant was chosen (1) to be enough to pull the conduction band edge close
to the Fermi level at the location of the ionized impurities and (2) to attain
the same doping level for all QWs.
Table 1.1: Essential structural and electronic properties of samples M242
and M280. Here, d
B
is the thickness of the barriers; other symbols are
dened in the text [see also Figs. (1.3), (1.4), and (1.5)]. The thicknesses (w
and d
B
) and the composition (x) were determined from calibrated growth
rates; w
0
is the rms width of the self-consistently calculated subband wave
function in each QW. The zero energy coincides with the bottom of the
QW. The energy levels E
0
QW
and E
1
QW
for sample M242 are results of self-
consistent calculations, whereas for sample M280 they are estimates based
on Eq. (1.3).
w w
0
d
B
E
0
QW
E
F
E
1
QW
Sample

A
x
meV
n [10
11
cm
2
]
M242 250 65 1850 0.3 10.5 15.5 31 1.40 0.02
M280 300 71 2500 0.1 6 9 25 0.86 0.02
While -doping technique was used to minimize the remote impurity scatter-
ing, sample quality also depends on the unintentionally incorporated impuri-
ties. Such impurities could be introduced by the substrate itself; subsequent
migration with the growth front results in contamination of the interfaces
were the 2DES is formed. To combat this an AlAs/GaAs gettering super-
lattice is grown rst. In such a superlattice, migrating impurities tend to
be trapped at the AlAs/GaAs or GaAs/AlAs interfaces. Important struc-
tural and electronic properties of M242 and M280 samples are summarized
in Table (1.1). Note that the measured density is essentially the same as the
nominal Si-doping. In other words, in these samples practically all electrons
from the donors are transferred to the QWs.
12 CHAPTER 1. THE QUANTUM HALL EFFECT
1.2.2 Electrons in GaAs quantum wells
The GaAs crystal has a zinc blende type of structure. Representative prop-
erties of bulk GaAs at room temperature are given in Table (1.2) according
to the work of Blakemore [12]. In GaAs, there are 8 valence electrons per
unit cell which contribute to the chemical bonds. The other electrons of each
kind of atom are frozen in closed shell congurations and they do not con-
tribute to the electronic properties investigated in this thesis. The valence
electrons hybridize to form tetrahedral bonds between one kind of atom (say
Ga) and its four nearest neighbors (As). In other words, the orbitals of every
atom (s-like or p-like) hybridize with the orbital of the neighboring atom,
thus producing two levels: one bonding and one antibonding. Along the
Ga-As bonding chain, the orbitals have most of the charge density shifted
towards As ion. The bonding s-orbitals are always occupied by 2 valence
electrons per unit cell. The remaining 6 valence electrons per unit cell oc-
cupy the three p-bonding orbitals. Because there is a large number of unit
Table 1.2: Unit cube size, nearest-neighbor distance between similar atoms,
atomic density, molecular weight, and crystal density of bulk GaAs [12].
Parameter Symbol Value
Length of side of unit cube A
c
5.653

A
Nearest-neighbor distance d
c
=

2A
c
/2 3.997

A
Atomic density 8/A
3
c
4.428 10
22
cm
3
Molecular weight M
c
145 amu
Crystal density
c
5.317 g cm
3
cells, bonding and antibonding levels broaden into bands. The top of the
valence band in GaAs occurs at the center of the Brillouin zone ( point). In
the absence of the spin-orbit coupling (see below), the three valence bands
(which originate from p-bonding orbitals) are degenerate at the point.
The bands originating from the antibonding orbitals are all empty, the low-
est lying s-band forming the conduction band of the material. In GaAs the
bottom of the conduction band occurs at the point and the bandgap is
about 1.52 eV at low T. In short, for GaAs, at the point of the con-
duction band, the s-like contribution dominates the total electron charge
density distribution, which has an antibonding character with most of the
1.2. ELECTRONIC PROPERTIES 13
charge localized on the As ion. In the following, I would like to remark on
some important properties of electron-doped GaAs QWs.
The eective mass and the eective Lande g-factor
The overall eects of the GaAs band structure, i.e., the fact that the electron
experiences the periodic potential of the lattice instead of moving in the
vacuum, are embodied in the use of an eective mass instead of the free
electron mass
6
. The relevant eective mass of 2D electrons in our samples
is assumed to be the isotropic conduction electron mass in bulk GaAs, m

=
0.067m
e
.
One detail still needs to be claried. What happens to the g-factor of
the electron? Up to now we considered the electron spin completely in-
ert dynamically, but in GaAs electrons move in an eective electric eld
that results from the built-in potential of the non-symmetric zinc-blende
structure of the underlying crystal. As a result, an electron experiences an
additional potential proportional to the scalar product of its spin magnetic
moment with the vector product of its velocity and the crystalline electric
eld. This additional interaction is referred to as the spin-orbit coupling
and it leads to a renormalization of the g-factor of the free electron. In bulk
GaAs, conduction electrons have an isotropic eective g-factor: g

= 0.44.
In GaAs QWs, the quantum connement renders the g-factor of 2D elec-
trons anisotropic and modies the values of the components of the g-factor
along (g
|
) and perpendicular (g

) to the growth axis of the heterostruc-


ture. Experimental studies in GaAs/Al
0.3
Ga
0.7
As systems revealed that for
w 120

A there is no observable g-factor anisotropy [72]. It has been
also veried that for w 200

A the g-factor is close to g

= 0.44 [72].
Therefore, in our samples, the eective g-factor of 2D electrons could be
considered isotropic and equal to g

= 0.44 in all experimental situations.


The electronic charge distribution
The problem of charge density distribution in GaAs QWs demands eluci-
dation. If one considers a 2DES in the presence of an external potential,
the spatial electronic distribution is described in general by the Schrodinger
equation for the wave function (containing the connement potential) and
6
The eective mass will replace the free electron mass in Eqs. (1.2), (1.3), (1.4), (1.5),
(1.6), (1.7), and (1.9). The constant DOS for 2DESs in GaAs QWs is m

/(
2
) = 2.8
10
10
cm
2
meV
1
.
14 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.5: The numerically calculated charge distribution
e
(z) (solid
curve) within the QW conning potential (dotted curve) for sample M242.
Open circles describe a Gaussian t to
e
(z), dening the rms width w
0
of
the 2DES. [Courtesy of S. Shukla]
the Poisson equation for the potential (containing the wave function through
the charge density). These have to be solved together self-consistently. The
realistic connement potential energy is represented by the conduction band
edge of the host material. The quantum mechanical description of sample
M242 is summarized in Fig. (1.5), which shows the part of the band diagram
containing the resulting conduction band prole after the charge transfer has
taken place
7
. Table (1.1) includes the eigenenergy solutions. The second
subband is unoccupied and energetically distant from the ground subband,
7
The modelling assumed a perfect symmetry with respect to the center of the QW and
employed the measured electron density.
1.2. ELECTRONIC PROPERTIES 15
and may be safely ignored. The ground subband wave function (
0
) has
a nite spatial extent in the z-direction. The nite thickness of the 2DES
could be simply estimated by representing the exact wave function in terms
of a Gaussian distribution of rms width w
0
[Table (1.1)]. However, in order
to simplify the calculations, it is usually assumed that the 2DES is ideal,
i.e., the 2DES is placed in a sheet of zero thickness, surrounded by an
homogeneous medium with a dielectric constant = 13. This value is close
to the dielectric constant of bulk GaAs.
1.2.3 The Landau level diagram
The eective Lande g-factor and the eective mass dictate the energy spec-
trum of 2DESs in the presence of an external magnetic eld. If we properly
include m

and g

into the description of 2DESs in GaAs QWs, the single-


particle Hamiltonian writes as
H
1e
=
1
2m

(p +eA)
2
+[g

[
B
s B. (1.12)
Its energy eigenvalues are given by
E
N,
=
c
(N + 1/2) +[g

[
B
B;
c
= eB

/m

. (1.13)
In the presence of a perpendicular magnetic eld, the electron energies are
labelled by two quantum numbers [the electric subband index (r) and the LL
index (N)] plus the spin variable =
1
2
. Each electric subband develops
into a series of spin-split LLs counted from the subband edge. Assuming
that only the ground electric subband is occupied, the DOS per unit area is
T(E) = n

N,
(E E
N,
), (1.14)
where n

= eB

/h is the degeneracy of one spin-split LL. The nominal


number of lled LLs at any magnetic eld is called LL lling factor , and
is given by = n/n

. We shall not extend the discussion of LLs other than


to remark that an accurate description of 2DESs in GaAs QWs at B ,= 0
should also take into account the correlation-exchange eects between elec-
trons. Exchange eects (which keep electrons with antiparallel spin apart)
were included into the diagram depicted in Fig. (1.6), which shows only the
LLs relevant for the experiments presented in this thesis. In the presence of
a perpendicular magnetic eld, one may talk about electric subbands sepa-
rated in energy by both exchange and Zeeman eects and then, further split
into LLs.
16 CHAPTER 1. THE QUANTUM HALL EFFECT
This point of view of two spin bands is particularly useful for understanding
several experimental results, and we shall adopt it here. In Fig. (1.6) the
energy gap between (N = 0, ) LL and (N = 0, ) LL is exchange enhanced
and it equals the sum of E
Z
and E
xc
, whereas the energy gap between
(N = 0, ) LL and (N = 1, ) LL is E

.
Figure 1.6: The energy spectrum of 2D electrons conned to GaAs QWs in
the presence of a quantizing perpendicular magnetic eld is schematically
shown. Orbital (E

), exchange (E
xc
), and Zeeman (E
Z
) eects are included.
1.3 Magnetotransport: Experimental details
All electrical resistance data discussed in this thesis were collected in an
Oxford Instruments Kelvinox 300 Dilution Refrigerator (Bayotron) with a
measured base temperature of 8 mK. The design of the base of the mixing
chamber, made in Stycast epoxy, is illustrated in Fig. (1.7). Samples are
mounted in vacuum by means of a (dual in-line multiple pin) DIP-socket
which is xed onto the mixing chamber very near to the copper cold n-
ger. At zero magnetic eld, while monitoring the RuO
2
mixing chamber
thermometer, typical temperatures that can be maintained continuously in
this apparatus are below 20 mK. The Bayotron is presently used with a
superconducting solenoid designed to provide a maximum central magnetic
eld of 15 T at 4.2 K and 17 T at 2.2 K. The spatial homogeneity of the
magnetic eld (over a 10 mm diameter spherical volume) is 1 part in 10
3
.
The dilution refrigerator is equipped with a tilting stage (see for details the
experimental section in Chapter 3) which can be rotated in situ, so the sam-
1.3. MAGNETOTRANSPORT: EXPERIMENTAL DETAILS 17
ples can be rotated from the horizontal to vertical (perpendicular to parallel
to the applied magnetic eld). The tilt angle is dened as being the an-
gle between the direction of the applied magnetic eld and the normal to
the sample plane. Unless otherwise stated, the magnetic eld was applied
perpendicular to the plane of the 2D electron layers.
The geometry of our measured specimens diers from the traditional
geometry employed for Hall resistivity measurements. The so-called van
der Pauw geometry, as illustrated in Fig. (1.1), has been used throughout
the present electrical transport experiments [106]. Each sample was cleaved
(along the natural cleavage directions 110) and 110)) into a small square
(approximately 2 to 3 mm on a side) from the MBE grown wafer. In the
standard procedure of fabricating ohmic contacts to the 2DES, the sample
was rst cleaned in trichlorethylene, acetone, methanol, and blown dry with
N
2
gas. Then it was held face up in a vacuum chunk and metal contacts
were deposited by hand with a soldering iron, under a microscope. Metal
contacts consisted of eutectic mixture of In : Sn ( 50% Sn) and were
apposed on the corners and at the centers of the edges of the sample
8
.
The sample and alloy contacts were then annealed in an oxide-reducing
hydrogen atmosphere N
2
: H
2
( 5% H
2
) at typically 440 450

C for 15-
25 minutes to form ohmic contacts to the buried electron layers. With care
under a microscope, the sample was then held in the vacuum chunk and
25 m-diameter gold wires, which serve as electrical leads, were soldered
to the ohmic contacts. The next step consisted of mounting the sample
on a DIP-header, by soldering the electrical leads to the pins with indium.
The DIP-header supporting the sample ts either into the standard DIP-
socket mounted on the tilting platform or into the DIP-socket xed onto the
mixing chamber [Fig. (1.7)]. The sample was placed in close proximity to the
calibrated RuO
2
chip resistor (with known magnetoresistance corrections)
employed as a thermometer. The thermometer was attached to the sample
through a copper strip, which was glued to the DIP-header with GE7031
varnish. An example of the nal result of this procedure is shown in Fig. (1.7)
for a sample cleaved from the M280 wafer.
In the context of magneto-transport measurements on multiple-QW sam-
ples we make two crucial remarks. First, the carrier concentration was not
changed through a persistent photoconductive eect in any of experiments
presented in this thesis
9
. Second, in order to ensure a good well-to-well ho-
8
It is very important that the metal for the contacts touches the edge of the sample to
prevent Corbino-geometry contacts.
9
Modulation-doped conventional heterointerfaces are usually sensitive to the exposure
of light. A red light emitting diode is commonly used to vary the electron density through
18 CHAPTER 1. THE QUANTUM HALL EFFECT
mogeneity, samples were cooled slowly (over several hours) from room tem-
perature to the liquid helium temperature. In such state-of-the-art multiple-
QW heterostructures we expect that the average electronic density varies by
only a few percents from layer to layer. In a typical specimen with maxi-
mum size 3 3 mm, the density uctuations across the sample are below
2%. Despite the fragility of the sought-after behavior, experimental reasons
dictated the use in magneto-transport experiments of samples cleaved near
the edge of MBE grown wafers. The quality of specimens coming from the
marginal parts of the wafer is slightly worse than those cleaved from the
central part. Furthermore, the density across the 2-inch GaAs wafer typ-
ically varies by 15 % from the nominal values calibrated near the wafer
center, resulting in a slightly larger density of samples cut from the edge of
the wafer than those taken from the central part.
For magneto-transport measurements a standard ac lock-in technique
has been used. By means of the ohmic contacts a constant low-amplitude,
low-frequency ac current is driven through the sample. The voltage in phase
is recorded, with a lock-in amplier across contacts at dierent positions.
Two measurements are made: the voltage drop in the direction parallel
to the current ow (V
L
) and the voltage dierence across the current ow
(V
H
). Voltages V
L
and V
H
were recorded and divided by the applied current
I, to obtain the longitudinal resistance (R
L
= V
L
/I) and the Hall resistance
(R
H
= V
H
/I). The longitudinal resistance could be optionally scaled to
obtain the longitudinal resistivity, while the Hall resistivity is exactly equal
to the measured Hall resistance
10
. In a rst series of experiments, the
transport coecients R
L
and R
H
are measured as a function of B, at xed T.
As our samples were not immersed in the
3
He/
4
He mixture, care was taken
to avoid heating of the 2DES above the bath temperature by the excitation
current. We used excitation currents of 1-100 nA rms at frequencies <
100 Hz, and the sweeping magnetic eld rate was kept below 0.1 T/min to
ensure the thermal equilibrium condition. The second series of experiments
performed at xed B, in which R
L
is measured as a function of T, will be
described later on.
a persistent photoconductive eect.
10
In two dimensions, the Hall resistivity
H
equals R
H
, independent of the sample
geometry. The longitudinal resistivity is
L
= (/ ln 2)R
L
for van der Pauw square-shaped
samples. We assume that an adequate annealing of ohmic contacts to all 100 buried 2D
electron layers was obtained for our measured specimens. Though the details of the current
path may be intricate, we consider that each multiple-QW structure behaves as a stack
of one hundred well-separated 2DESs (resistors) connected in parallel. Accordingly, the
longitudinal resistance is either given for all electrically contacted 2D electron layers (as
measured) or quoted per single layer. The Hall resistance is always given per one layer.
1.3. MAGNETOTRANSPORT: EXPERIMENTAL DETAILS 19
Figure 1.7: (Top) View of the Bayotron mixing chamber. (Bottom) Exper-
imental conguration for magneto-transport measurements (sample M280).
20 CHAPTER 1. THE QUANTUM HALL EFFECT
1.4 Fundamental aspects of the QHE
1.4.1 Measuring the QHE
To go any further toward understanding the QHE it is useful to briey look
now at the ordinary Hall eect, discovered in 1879. In the Drude theory
of the electrical resistivity of a simple metal, an electron is accelerated by
the applied electric eld for an average time
0
, the mean free time (alias
the transport scattering time), before being scattered to a state which has
average velocity zero. In the free electron model, the electrical resistivity
at zero magnetic eld is
0
= m

/(ne
2

0
). The carrier mobility at zero
magnetic eld is traditionally dened as
0
= (ne
0
)
1
. In the presence of
the magnetic eld the electrons path is curved, due to the Lorentz force,
and the resistivity tensor writes as
_

L
B/(ne)
B/(ne)
L
_
(1.15)
The longitudinal resistivity
L
has the same signicance as the conventional
notion of the electrical resistivity of an elemental material. The Hall resis-
tivity is zero in the absence of a magnetic eld and increases linearly with
magnetic eld according to
H
= B/(ne). These expressions have been
extensively compared to the experiment and they are often good approx-
imations. These simple results have been obtained within a semiclassical
theory with quantum mechanics entering very indirectly through the values
of
0
and m

. Only few decades ago they were expected to remain valid at


low-T for 2DESs in the presence of very high magnetic elds. The pioneer-
ing paper of Fowler et al. [37] was the rst to establish experimentally the
existence of quantum corrections to the longitudinal resistivity of a 2DES in
a perpendicular magnetic eld. These corrections produce the well known
Shubnikov-de Haas (SdH) oscillations.
Lets turn now our attention to low-T magneto-transport experiments
on samples M280 and M242. Figure (1.8) displays a detailed synopsis of
the evolution of magnetoresistance with decreasing T in sample M280. Fig-
ures (1.9) and (1.10) show low-T traces of the transport coecients R
L
and
R
H
, in magnetic elds up to 15 T, for samples M280 and M242, respec-
tively. The detailed structure observed for transport coecients in M280
sample, at low magnetic elds (B 1 T), is presented in Fig. (1.11). Con-
centrating on the low-B regime (B 0.2 T) [Fig. (1.11)], we rst note that
the Hall resistance is proportional to the magnetic eld, same as in the
classical Hall eect. The longitudinal resistance presents SdH oscillations,
1.4. FUNDAMENTAL ASPECTS OF THE QHE 21
with the characteristic periodicity in 1/B

, as expected. The SdH struc-


ture becomes stronger with decreasing T. We have observed only one set
of oscillations, indicative of one subband of 2D electrons. As B is ramped
up, under the condition of Landau quantization, E
F
moves through succes-
sively LLs. The period in 1/B

of magnetoresistance oscillations is given


by e/(m

E
F
) = (2/n)(e/h). The electron density determined from the
position of R
L
minima (h/e = 4.137 10
11
T cm
2
) agrees with the density
deduced from the slope of R
H
. By assigning the lling factor according to
= 2l
2
B
n, one can explicitly label the minima displayed by the longitudinal
resistance. At the lowest investigated temperatures, the LL spin-splitting is
readily visible at = 15 in sample M280. In the present experiments, the
zero magnetic eld electrical resistivity is T-independent below 2 K. The
estimated low-T zero magnetic eld electron mobility (
0
) in our samples is
given in Table (1.3).
Figure 1.8: Overview of the B dependence of the longitudinal resistance R
L
at = 0

and indicated temperatures (sample M280). Note that the widths


and depths of the R
L
minima dramatically increase for T 0.
22 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.9: Longitudinal and Hall resistance [R
L
(left axis) and R
H
(right
axis)] vs magnetic eld at T = 50 mK and = 0

(sample M280). The


vertical lines indicate some prominent integral and fractional lling factors.
1.4. FUNDAMENTAL ASPECTS OF THE QHE 23
Figure 1.10: Longitudinal and Hall resistance [R
L
(right axis) and R
H
(left
axis)] vs magnetic eld at = 0

and indicated temperatures (sample M242).


The vertical lines indicate some prominent integral and fractional lling
factors where the QHE is observed.
24 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.11: Longitudinal and Hall resistance [R
L
(left axis) and R
H
(right
axis)] vs magnetic eld at T = 50 mK and = 0

revealing the Shubnikov-de


Haas [SdH] regime (B 0.3 T) and the quantum Hall eect [QHE] regime
(B 0.3 T), and the spin-splitting collapse at = 15 (sample M280). The
dashed line indicates the expected classical behavior for the Hall resistivity.
1.4. FUNDAMENTAL ASPECTS OF THE QHE 25
Lets move now to the high magnetic eld region. Figures (1.9) and (1.10)
display two astonishing aspects which form the hallmark of the QHEs. First,
the longitudinal resistance fall essentially to zero over wide ranges of B. For
example, measurements of the T-dependence of R
L
at = 1/3 (B 11 T)
on M280 sample clearly show the development of a zero-resistance state
[Figs. (1.8) and (1.10)]. The second aspect, even more astonishing than the
rst, is that concurrent with the vanishing R
L
, there are plateaus in the
Hall resistance. Note that with decreasing T, the B-ranges over which the
longitudinal resistance is essentially zero become larger and larger, whereas
the width of the plateaus tends to its maximal value, which corresponds
to perfectly abrupt steps in R
H
. Close examination of the values of R
H
at these plateaus reveals that all can be described by a universal formula
R
H
= h/(e
2
), with either an integer or a simple rational fraction with
odd denominator. The richness of the phenomenon is evidenced by the
large number of observed fractions
11
. Another important feature of the
experimental data is that the higher the
0
, the more prominent is the
fractional QHE. This point is very clear from the data shown in Fig. (1.9)
(high-mobility sample) and Fig. (1.10) (low-mobility sample).
1.4.2 Integral QHE
Based on phenomenological grounds, it seems clear that what is required for
the observation of QHEs is an energy gap , separating the ground state
and its current carrying excitations. In this context, one important aspect to
be established is the criterion for the delineation of the quantum-transport
regime, separating it from the classical regime. In general, two conditions
should be fullled simultaneously for the macroscopic observation of quan-
tum eects in 2DESs: and k
B
T . Here, = /
s
(in units of
energy) is associated to the single-particle scattering time
12
. Since mea-
surements are typically performed at dilution refrigerator temperatures, the
11
Note the weak resolution of the R
L
minima at fractions in sample M242. This is
attributed to the poorer quality of the sample used in the present experiments compared
to the previously investigated sample [8] (cleaved near the center of the MBE grown wafer).
12
In the transport theory of 2DESs one must deal with two dierent characteristic times
- a single-particle scattering time
s
and a transport scattering time
0
. The single-particle
scattering time is a measure of the time for which an electronic momentum eigenstate can
be dened in the presence of scattering. The transport scattering time is given by
0
and
could be related to a disorder broadening parameter
0
which is an informative measure
of the degree of disorder in real samples. For the 2DESs studied here, estimates of
0
,
based on the expression
0
= /
0
, are of the order of 1 K. We note that
0
could be two
orders of magnitude larger than
s
[25].
26 CHAPTER 1. THE QUANTUM HALL EFFECT
condition k
B
T is trivially satised. As the presence of a quantizing
magnetic eld is a necessary premise for the observation of QHEs, we ex-
pect that is in some way related to B. In fact, we have already met in
the single-particle energy level calculations, deguised either in the cyclotron
splitting or in the Zeeman splitting. One of the most transparent illustra-
tions of this general criterion for macroscopic quantum behavior in 2DESs
can be found in the occurrence of SdH oscillations; magneto-quantum oscil-
lations in R
L
will appear when the magnetic eld is large enough such as
E

. For even larger magnetic elds, the Zeeman splitting will become
resolved when E
Z
.
In combination with the single-particle quantization eects, disorder is
the essential ingredient for observing the integer QHE. In the most simple
theory of the integer QHE, one follows the one-electron analysis and looks at
the eects of disorder on the energy spectrum of 2D electrons in a quantizing
magnetic eld
13
. In a real 2DES, LLs are broadened into a set of energy
levels due to the nite amount of the disorder present in the sample; the
DOS consists of a series of non-ideal -functions centered at the discrete
LL energies [Fig. (1.12)]. For mathematical convenience, in the present work,
the shape of each LL is considered to be a Gaussian [114]. The rms width of
each LL is denoted by
G
. (The full width at the half maximum is 2.36
G
and may be compared to
0
= /
0
.) If only the ground electric subband is
occupied, the Gaussian DOS is
T
G
= n

N,
(2)
1/2

1
G
exp (E E
N,
)
2
/(2
2
G
). (1.16)
More generally, the DOS could be obtained from a self-consistent theory
where the impurity-induced level broadening and screening determine each
other [68]. Specically, when E
F
lies at the middle (edge) of a Landau level
screening is strong (weak) and the screened potential is short ranged (long
ranged). It should be noted that self-consistent DOS calculations become
intractably complex when many LLs are occupied.
13
There is a glaring discrepancy between von Klitzings discovery and the localization
theory [E. Abrahams, P.W. Anderson, D.C. Licciardello, and T.V. Ramakrishnan, Scaling
Theory of Localization: Absence of Quantum Diusion in Two Dimensions, Phys. Rev.
Lett. 42, 673 (1979).] Localisation theory predicts that weak disorder is sucient to
localize non-interacting electrons in 2D at B = 0 and zero temperature. The occurrence
of QHEs, which necessarily implies the existence of delocalized states, gave new impetus
to the study of the 2D metal-to-insulator transition at B = 0.
1.4. FUNDAMENTAL ASPECTS OF THE QHE 27
The crucial point is that some states within the broadened LLs will be lo-
calized and the others will be extended. It turns out that there is exactly
one extended state per LL, so that the lling factor equals the number of
occupied extended states. These extended states are located at E
N,
. The
basic picture of the integral QHE is as follows. The Hall resistance is quan-
tized whenever E
F
lies in a mobility gap, i.e., a region of localized states
14
.
When E
F
= E
N,
, the Hall resistance jumps from one quantized plateau to
the next
15
. As long as E
F
moves through a region of localized states, the
longitudinal resistance keeps its value, which is essentially zero. The peaks
observed in the R
L
curves correspond to maximal electron scattering, which
occurs when E
F
coincides to E
N,
.
One may succinctly summarize the integral QHE in the following
way [27]. As E
F
passes through the critical energy E
N,
there is a
insulator-metal-insulator transition (at T = 0), with the 2DES metal-
lic precisely at E
F
= E
N,
. The insulator-metal-insulator transition can
be understood as a percolation transition. As the percolation level is ap-
proached, the edge states on the two sides of the sample will begin taking
detours deeper and deeper into the bulk and begin to communicate with
each other. When the percolation transition occurs edge states become part
of the bulk and the sample is a normal metal.
Finally, we call attention to the fact that in transport eects the local-
ized and the extended states are both of crucial meaning. For the equi-
librium properties, such as specic heat, the dierence between localised
and extended states is of no importance, i.e., equilibrium properties depend
primarily on the total DOS [26, 70].
14
A short note on the edge states is in order here. To understand edges it is necessary
to realize that they are normal metals with dissipation. In a real sample, the mobility
gap collapses in a complicated way near the edge of the sample, so as to make the edge a
normal metal with a well dened chemical potential. The detailed nature of the edge is
unimportant, as long as the net result of a QHE gedanken experiment, in which the 2DES
is bent into a loop, is the transfer of electrons from the local Fermi level at one edge to
the local Fermi level at the other, without dissipating energy. Laughlins gauge argument
for the integral QHE [66] relates the quantized Hall resistance to the charge of electrons
transferred in a gedanken experiment.
15
The dependence of the plateau width upon the electronic mobility suggests that QHE
plateaus originate from a mobility gap.
28 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.12: Single-particle DOS vs energy of a 2DES in a quantizing mag-
netic eld. (Top) Disorder free 2DES. (Bottom) Realistic 2DES. The shaded
regions consist of localized states. Gaussian broadening of LLs was assumed.
1.4.3 Fractional QHE
What is common to the integer and the fractional QHE is the formation of
plateaus in the Hall resistance concomitantly with a vanishing longitudinal
resistance as T 0. The fact that R
L
has an activated behavior with T in
the plateau region indicates the existence of a nite excitation gap between
the QHE ground state and its excited states. The explanation of the frac-
tional QHE can not be sustained without basing it upon the existence of
a many-particle excitation gap, whose physical origin diers fundamentally
from that of the integer QHE excitation gap. Previously, in explaining the
integer QHE, we started with a single-particle Hamiltonian. Building upon
1.4. FUNDAMENTAL ASPECTS OF THE QHE 29
the non-interacting electron picture, we considered the eect of the disorder
potential on the energy spectrum. With the formation of disorder broadened
(spin-split) LLs we arrested the necessary (single-particle) excitation gap.
Then, we argued that whenever the chemical potential lies within a mobility
gap which occurs at integer lling factor = i, the Hall resistance depends
only on i through h/(ie
2
). At low lling factors the 2DES enters a regime
where, in the absence of disorder, its ground state is determined entirely
by electron-electron interactions
16
. For < 1 the meaning of this is clear
enough: the kinetic energy is an irrelevant constant and the Hamiltonian
has only one energy scale set by the Coulomb interaction
17
. Therefore,
for understanding the fractional QHE one should rst consider the eect of
interactions between electrons. The many-particle Hamiltonian relevant for
the fractional QHE was given by Laughlin in his celebrated paper on the
Anomalous Quantum Hall Eect: An Incompressible Liquid with Fraction-
ally Charged Excitations [67]. With a priori unjustiable omission of the
Zeeman term, it writes as
H
FQHE
=
1
2m

i=1
_
p
i
+eA
i
(r
i
)

2
+
1
2
A

i,=j=1
V
ee
(r
i
r
j
), (1.17)
where ^ is the number of electrons, the potential is the Coulomb repulsion
energy V
ee
(r) = e
2
/([r[) (in CGS), and the disorder, which would enter
through an additional potential energy term, is ignored. Laughlin proposed
a many-electron (spin-polarized) variational wave function of the form

L
m
(z
1
, z
2
, . . .) =
L
m
[z] = P
m
[z]
A

k=1
e

1
4
[z
k
[
2
=
A

i<j
(z
i
z
j
)
m
A

k=1
e

1
4
[z
k
[
2
.
(1.18)
The Laughlin wave function and plasma analogy formalism describe stable
states of the 2DES, for which m is related to the lling factor through
= 1/m. The value of m in Eq. (1.18) can be deduced by noticing that P
m
[z]
must change sign when any pair of particles are interchanged. The Pauli
16
The strength of the disorder potential in our samples is much smaller than E
C
. Only
such low-disorder (clean) samples are promising hunting grounds for the fractional QHE.
17
Coulomb repulsion between electrons causes them to condense into highly correlated
many-body ground states. The sequence of fractional QHE ground states terminates in a
insulating state (presumably a Wigner crystal).
30 CHAPTER 1. THE QUANTUM HALL EFFECT
exclusion principle requires that m is a positive odd integer (m = 1, 3, 5, . . .),
and since = 1/m, the fractional QHE may occur at odd-denominator
lling factors ( = 1/3, 1/5, . . .). Thus, the most striking feature of the
theory is the implication that, for an ideal, spinless 2DES with no impurity
scattering, there exists at T = 0 a sequence of fractional QHE ground states
at = 1/m, with m an odd integer. Experimentally, the fractional QHE
occurs at a large number of other lling factors. Theoretical studies have
shown that the = 2/3 fractional QHE state is an electronic conjugate of
the = 1/3 fractional QHE state. Invoking the particle-hole symmetry in
the lowest LL, the state at = 1 1/m is composed of holes behaving
as electrons at = 1/m. Furthermore, with the inclusion of spin, the
particle-hole symmetry dictates that fractional QHE states at and 2
are identical. Thus, the fractional QHE state at = 4/3 is expected to
have the same phenomenology as the = 2/3 fractional QHE state. Note
that all prominent fractional quantum numbers (less than unity) observed
in our samples consists of two hierarchies of continued-fraction lling factors
derived from 1/3 and 2/3
i
2i + 1
=
1
3
,
2
5
, . . .
i
2i 1
=
2
3
,
3
5
, . . . ,
where i is a positive integer.
By the internal logic of his argument, the logic of the trial wave function,
Laughlin was led to the conclusion that any elementary charged excitation in
the fractional QHE has to carry a fractional charge and lies at a nite energy
above the many-body ground state
18
. The quasiparticle bands, separated in
the disorder-free case by the (many-particle) excitation gap

, are broad-
ened in the presence of disorder into a continuum consisting of two bands
of extended states separated by a band of localized ones. For the fractional
QHE state at = 1/m, the Hall resistivity is related by gauge invariance to
the charge of the quasiparticles e

= e/m through R
H
= h/(ee

). Whenever
the chemical potential lies in the localized state quasiparticle band, R
H
is
quantized to R
H
= h/[(1/m)e
2
]. This result remains valid for arbitrary frac-
tional quantum numbers, i.e., when 1/m p/q. The theoretical excitation
18
What is necessary to explain the fractional QHE is a downward cusp in the total energy
of the system at particular fractional lling factors. The existence of a cusp in the total
energy implies a discontinuity in the chemical potential, which implies, in turn, that there
is a gap in the spectrum of charge carrying excitations. At densities where the chemical
potential has a discontinuity the system is said to be incompressible.
1.4. FUNDAMENTAL ASPECTS OF THE QHE 31
gap for a pair of one free quasielectron and one free quasihole at = p/q is


= C
q
e
2
l
B
. (1.19)
This should be compared with the QHE excitation gap determined in trans-
port experiments
19
. The prefactor C
q
depends on the denominator q and is
model dependent. For = 1/3, 2/3, 4/3, . . . theories have been converging
toward a value of C
3
0.05 0.1, whereas for = 2/5 or 3/5 theoretical
estimates give C
5
0.015 0.030. Estimates for the QHE excitation gaps
(
t

) at prominent fractions in sample M280 are given in Table (1.5).


The construction of Laughlin quasiparticles is very easy to visualize and
predicts a remarkably successful phenomenology. Imagine piercing the in-
compressible 2DES at lling factor = 1/m with a innitely thin solenoid
located at a point we take to be the origin, as illustrated in Fig. (1.13).
Then, the ux through the solenoid is adiabatically changed from 0 to
0
.
Once an entire ux quantum has been added, the Hamiltonian has, up to
a gauge transformation, evolved back to its value at zero ux. Thus, the
solenoid threaded by
0
may be gauged away; the state generated by this
process (if the ground state has no degeneracy) is an exact excited state of
the original Hamiltonian. Furthermore, one can show that this state has
charge e/m added to an area surrounding the origin. When magnetic ux is
added through a disk of radius R of the system, an azimuthal electric eld
E

(t) =
1
2R
d
dt
is generated by virtue of Faradays law. This electric eld
drives a radial current j
r
(t) =
H
E

(t) =
1
m
e
2
h
d
(2R)dt
, where
H
=
1
m
e
2
h
is the quantized Hall conductance. As the time derivative of the charge
accumulated inside the disk is
_
j
r
(t)dr, the total charge added for one
ux quantum increase is e/m. Since electric neutrality has to be preserved
a charge e/m also appears outside the disk. To be precise, threading one
ux quantum generates one quasiparticle plus one quasihole. The charge of
the quasihole is e/m. The change in total energy when one ux quantum
19
Thermodynamic (

) and transport (

) excitation gaps are related by

= m

at lling fraction = 1/m. This follows from the fractionally charged nature of the quasi-
particle/quasihole pairs which control the current response of the 2DES at fractional lling
factors. Since the addition of a real electron to the = 1/m state generates m quasiparti-
cles, thermodynamic gaps are m times larger than transport gaps. Thermodynamic gaps
are usually inferred from the measured magnetization jumps [112]. As soon as the 2DES
has two compressible regions (alias the edges), separated by a incompressible region per-
colating through the sample, a deep minimum is observed in the longitudinal resistance.
The change in the potential across the incompressible region is

, which is related to
the magnetization jump by the relation M

= A

/
0
. For samples investigated in
this thesis, we expect that 70% of the sample area is incompressible at = 1/3.
32 CHAPTER 1. THE QUANTUM HALL EFFECT
is added to the system is 1/m times smaller than the energy necessary to
remove an electron.
Figure 1.13: Illustration of the Laughlins gedanken experiment leading to
fractional charge quasiparticle states at fractional LL lling factor = 1/3.
We are now in a position to consider (at a qualitative level) the nature of col-
lective modes in the QHE. Collective excitations of the QHE states related
to the charge degree of freedom are represented by density excitations. In
these excitations a particle (or quasiparticle) is promoted to an empty state
leaving behind an oppositely charged hole (or quasihole). For a wavevector
q there is a displacement between the centers of the two cyclotron orbits
that represent the single-particle states, given by l
q
= ql
2
B
. Collective modes
(labelled by q) are constructed with such a pair as a basis. The collec-
tive excitations can be regarded thus, as magnetic excitons with an average
dipole length l
q
. The inter-Landau-level excitations involve the transfer of
electrons from one LL to next higher LL, whereas in intra-Landau-level ex-
citations, electrons do not change Landau level quantum number. These
intra-Landau-level excitations reside lower in energy and are of principal
interest in the fractional QHE.
1.5. INTERLUDE: GROUND STATE THEORY AT = 1 33
1.5 Interlude: Ground state theory at = 1
If we digress now, it is in order not to lose information. We list the relevant
energy scales for the QHE in Table (1.3). In vacuum, neglecting the quantum
electrodynamics corrections, E
Z
= E

, but this is no longer valid for 2D


electrons in GaAs QWs. The cyclotron splitting, due to the small eective
mass of electrons in GaAs, is increased by a factor of m
e
/m

14, whereas
the applied magnetic eld scarcely couples to the spins due to the strong
spin-orbit interaction which reduces the electron g-factor by g
e
/[g

[ 5
times. The resulting Zeeman splitting, is thus some 70 times smaller than
the cyclotron splitting. There is a net decoupling of the scales of orbital
and spin energies. At low temperatures, it is possible for the 2DES to be
in a regime where the orbital motion is fully quantized (k
B
T
c
) but
the low-energy spin uctuations are not completely frozen out (E
Z
k
B
T).
The low energy spin dynamics of the 2DES is particularly interesting at
= 1. Since the Coulomb energy scale dominates at = 1, the system
spontaneously polarizes at T = 0 to minimize its energy. The 2D electron
spin polarization is complete as the kinetic energy has been quenched by the
magnetic eld
20
.
The 2DES at = 1 is our incompressible quantum uid prototype.
Clearly, itinerant ferromagnetism turns out to be a fundamental issue at
= 1 and we shall have occasion later on to study it more carefully
21
.
Here we intend to complete the specication of the Laughlin state at = 1
by inclusion of the spin degree of freedom. While the Coulomb force is
spin independent, exchange eects lead to additional complications in
the QHE. Pauli exclusion principle requires that the total many-body wave
function [z], which is the product between the orbital
L
1
[z] and the spin
part [
s
), changes sign under the simultaneous interchange of both space
20
The reader should be persuaded that there is no kinetic energy cost associated with
this electron spin polarization. The situation is dierent for a conventional ferromagnet
or Fermi gas of electrons. Please think at the criterion for a fully polarized ground state
in a 3D homogeneous electron gas at zero magnetic eld. The ground state energy for
ferromagnetic alignment of electron spins is
A

3
5
2
2/3
E
F

3
4
2
1/3
(e
2
k
F
)

.
The rst term is the kinetic energy and the second term represents the Coulomb exchange
energy. Only once the exchange term dominates the kinetic term the ferromagnetic ground
state is self-consistent.
21
The electron spin stiness, which reects the ordering of the electron spins, is an
energy scale closely related to E
C
(see Chapter 2).
34 CHAPTER 1. THE QUANTUM HALL EFFECT
and spin coordinates. At = 1 the symmetric spin part is
[
s
) = [
1

2
. . .
A
), (1.20)
and the antisymmetric orbital part writes as

L
1
[z] =
A

i<j
(z
i
z
j
)
A

k=1
e

1
4
[z
k
[
2
. (1.21)
Table 1.3: Estimates of relevant energy scales for samples M242 and M280
are given. They are based on values of the perpendicular magnetic eld
component at Landau level lling factor = 1 [5.9 T for M242 sample and
3.6 T for M280 sample] and electron mobility
0
values at zero magnetic
eld [3 10
5
cm
2
V
1
s
1
for M242 sample and 7 10
5
cm
2
V
1
s
1
for M280 sample]. Once the dierent energy scales have been identied, it
is important to convert them to the same units in order to compare them
easily. Here, all energies are given in K, [g

[ = 0.44 is the eective g-factor of


electrons in bulk GaAs, m

is the eective mass of electrons in bulk GaAs,

B
is the Bohr magneton, is the dielectric constant in bulk GaAs, B

is the
perpendicular component of the applied magnetic eld B, l
B
= [/(eB

)]
1/2
is the magnetic length, and
c
= eB

/m

. Rough estimates of important


dimensionless parameters are also given.
Energy Scale Expression M242 M280
Zeeman energy
a
E
Z
= [g

[
B
B 1.8 1.1
Coulomb energy
b
E
C
= e
2
/(l
B
) 121 95
Electron spin stiness
s
= e
2
/(16

2l
B
) 3 2
Cyclotron energy
c
E

=
c
118 72
Disorder broadening
0
= /
0
0.7 0.3
r
s
E
C
/E

1 1
g E
Z
/E
C
0.01 0.01
a
Expressed in K, E
Z
0.3 B [T].
b
Expressed in K, E
C
50 (B

[T])
1/2
.
c
Expressed in K, E

20 B

[T].
1.6. SUPPLEMENTAL ASPECTS OF THE QHE 35
Remarkably, at = 1, the Coulomb exchange energy per particle can be
calculated exactly. It equals (/8)
1/2
E
C
. Consequently, a quasiparticle-
quasihole Laughlin pair at = 1 causes a loss of exchange energy equal
to E
xc
= (/2)
1/2
E
C
. However, as spin and charge degrees of freedom are
connected at = 1, we need a more sophisticated mathematical language
to describe the lowest energy charged excitations in the system
22
.
1.6 Supplemental aspects of the QHE
1.6.1 Finite thickness corrections
In our discussion thus far, we have ignored the nite spread of the electron
wave function perpendicular to the 2DES plane. One of the most common
approximate forms assumed for the calculated charge prole in a QW is
the Gaussian distribution [
0
(z)[
2
= (2)
1/2
w
1
0
exp [z
2
/(2w
2
0
)], where
w
0
is the rms width. In our samples, w
0
is about a factor of 2 smaller
than l
B
at = 1. When w
0
is comparable to the magnetic length, the nite
thickness corrections are expected to have a substantial eect on the physical
quantities of interest, such as the QHE excitation gap or the ground-state
energy. The nite thickness of the 2DES softens the short-range divergence
of the bare Coulomb interaction and hence, reduces both the cohesive energy
of the ground state and the QHE excitation gap. In order to ascertain the
detrimental inuence of the nite spread of the electron wave function on the
QHE excitation gap and ground-state energy, the bare Coulomb interaction
is replaced by a more realistic force law of the form [23] where the eect of
the density prole in the z-direction is integrated out
V
w
(r
2D
) =
_ _

e
(z
2
1
+z
2
2
)/2w
2
0
2w
2
0
1
_
r
2
2D
+ (z
1
z
2
)
2
dz
1
dz
2
. (1.22)
Here, r
2D
is the 2D vector separating the two electrons, bare Coulomb in-
teractions correspond to w
0
= 0, and the energy is expressed in units of E
C
.
While the reduction of the ground-state energy still leaves the incompress-
ible quantum-uid state energetically favorable over the crystalline state,
the reduction of the QHE excitation gap is important for comparison with
the experiment [17]. Theoretical results for the QHE excitation gap at = 1
22
Note that, when only the charge degree of freedom is taken into account, the dispersion
relation for the collective neutral mode at = 1 for large q is E(q) = E

+E
xc
e
2
/(ql
2
B
).
At small q, E(q) = E

+ e
2
q/2.
36 CHAPTER 1. THE QUANTUM HALL EFFECT
in sample M280, which take into account the nite thickness of the 2DES,
will be presented in the next chapter.
1.6.2 Landau level mixing
The fractional QHE is observed at relatively large B and believed to occur in
the limit B . In this limit, since the Coulomb energy ( B
1/2
) is small
compared to the cyclotron energy ( B), the LL mixing could be safely
neglected. The degree of LL mixing can be expressed by the dimensionless
parameter r
s
, which is the ratio of the Coulomb energy to the cyclotron
energy. As r
s
B
1/2
, the degree of the LL mixing raises by decreasing
the magnetic eld. In order to make the discussion more quantitative and
for further reference, we note that in our samples r
s
at = 1 is of the
order of unity (r
s
1). Clearly, r
s
can not act as a small parameter for a
perturbation expansion. The LL mixing is critical for real 2DESs and must
be incorporated into realistic theories.
In the presence of LL mixing several aspects of the fractional QHE be-
come questionable. Among the most worrisome are the existence of the
fractional charge, the validity of the particle-hole symmetry, and the mag-
nitude of the QHE excitation gap. Laughlin oered a lucid explanation for
the exactness of the fractional quantization in the presence of extreme LL
mixing. The failure of the particle-hole symmetry between the fractional
QHE states at and 1 due to LL mixing, was rst investigated by Yosh-
ioka [18]. He computed the QHE excitation gaps at = 1/3 and 2/3 and
found a weak eect ( 10%), which is rather inconsistent with the large
asymmetry (
1/3
/
2/3
3) usually observed in experiments.
A real breakthrough in the subject came with Monte Carlo studies [76],
which calculate the QHE excitation gap at various lling factors by taking
into account both the LL mixing and the nite thickness of the 2DES. (We
shall tackle the case of lling factor unity in Chapter 2). Results obtained at
= 1/3 by Melik-Alaverdian et al. [76] show that the LL mixing has a weaker
eect on the spin-reversed quasiparticle excitation gap but has a strong eect
on the spin-polarized quasiparticle excitation gap. Interestingly, when one
rst includes the nite thickness of the 2DES, the additional correction due
to the LL mixing is essentially the same for both spin-polarized and spin-
reversed quasiparticle excitation gaps.
1.6. SUPPLEMENTAL ASPECTS OF THE QHE 37
1.6.3 Spin and the fractional QHE
Now how did the electron spin and the fractional QHE come to be con-
nected? In the limit of large Zeeman energy, all electrons have their spins
aligned with the magnetic eld and the spin degree of freedom is irrelevant
for the fractional QHE. Halperin was the rst to point out, in a prescient
paper [48], that for 2DESs in GaAs QWs, in most of the experimental cases,
E
Z
E
C
, E

and hence, it makes sense to re-examine the assumption of


complete electron spin polarization in the lowest LL. He went on to pro-
pose a family of generalized Laughlin wave functions that could incorporate
reversed spins. The spin-unpolarized QHE state at = 2/(m+p) writes as

m,m

,p
[z] =

i<j
(z
i
z
j
)
m

<
(z

)
m

(z
j
z

)
p

j
e

1
4
[z
j
[
2

1
4
[z

[
2
,
(1.23)
where Latin and Greek indices correspond to electrons with dierent spin
states. This wave function is characterized by the exponents m, m
t
(positive
integers) and p (non-negative integer) describing the relative angular mo-
mentum between species of the same spin and dierent spin, respectively.
In particular, a completely unpolarized wave function was constructed at
= 2/5 (m = 3, m
t
= 3, p = 2). Considering the spin-up electrons
and spin-down electrons as two dierent species of particles and using a
two-component classical plasma approach, Chakraborthy and Zhang [16]
calculated the ground state energy for the two-spin fractional QHE state
at = 2/5. With this result a very intriguing possibility to observe a spin-
reversed fractional QHE ground state was established. It indicated that, al-
though in the limit of large Zeeman energy the fully polarized ground state
is favored energetically, there always exists a possibility for a transition to
a not fully polarized state at lower Zeeman energy.
In addition to the ground-state spin conguration, spin should play a
role in determining the spectrum of excited states in QHE systems. For
instance, it was predicted that, while the = 1/3 QHE ground state is spin
polarized even for vanishing Zeeman energy, its lowest-energy elementary
excitations are spin-reversed quasiparticles at low B [17]. Theorists also
inquired how the QHE excitation gap will vary with the total magnetic
eld. For example, at low B, the lowest-energy excitations at = 1/3
were identied to be pairs of spin-reversed quasielectrons and spin-polarized
quasiholes. Accordingly, the QHE excitation gap rises linearly because of the
dominant Zeeman energy contribution. As B is increased further, a crossover
38 CHAPTER 1. THE QUANTUM HALL EFFECT
point is reached beyond which a fully polarized quasihole quasielectron pair
is energetically favored, and a B
1/2
dependence is obtained.
We hasten to point out that the standard approach used in magneto-
transport experiments to gain access to the spin degree of freedom is the
tilted magnetic eld technique, introduced by Fang and Stiles in their study
of the g-factor in Si inversion layers [35] and rst applied to the QHE by
Haug et al. [49]. In essence, the QHE excitation gap

is measured at xed
for various tilt angles. A straightforward interpretation of the results is
oered if we write

=
0

[g

[
B
BS, (1.24)
where S (in units of ) is the dierence in spin of the quantum uid state
before and after the excitation of the quasielectron-quasihole pair. It is
assumed that the total magnetic eld couples to the spin degrees of freedom,
while the perpendicular magnetic eld component B

controls the orbital


dynamics. The term
0

is the contribution to the gap from all non-Zeeman


sources and, in this model, depends only upon B

. Thus, the quantity


0

which is in general unknown, is an irrelevant constant in QHE excitation


gap studies at xed as a function of the tilt angle.
Table 1.4: Zeeman energy contribution to the excitation gap of a spin-
polarized QHE liquid [ . . . ) for various combinations of one quasielectron
and one quasihole with dierent spin-
1
2
polarizations.
Quasihole Quasielectron S
(1) () spin polarized () spin polarized 0
(2) () spin polarized () spin reversed -1
(3) () spin reversed () spin polarized -1
(4) () spin reversed () spin reversed -2
Lets examine the tilted magnetic eld dependence of the QHE excitation
gap in some detail, conning ourselves to the case when the quasiparticles
are fundamentally spin-
1
2
objects [Table (1.4)]. One can distinguish four
cases: (1) both the quasielectron and quasihole spins are polarized just like
the parent uid, (2) the quasielectron spin is reversed but the quasihole
remains polarized, (3) reverse of case 2, and (4) both quasiparticle spins are
reversed. For a spin-polarized QHE liquid, S = 0 for case 1, S = 1
1.6. SUPPLEMENTAL ASPECTS OF THE QHE 39
for both cases 2 and 3, and S = 2 for case 4, implying that

remains
the same or it may increase when the tilt angle is increased. On the other
hand, if the QHE liquid is partially spin-polarized it is possible for the net
spin to increases upon quasielectron quasihole pair excitation (S > 0).
In this case

will decrease when the tilt angle is increased, and eventually


vanishes. The above discussion is well-adapted to the purposes of this thesis;
it will be of great service for the case at hand of tilted magnetic eld QHE
excitation gap measurements.
1.6.4 QHEs in tilted magnetic elds
The assumption that the parallel component of the magnetic eld enters the
problem only through the Zeeman energy term is valid only if the 2DES is
innitely thin. Surely, in real experimental situations, the 2DES has a nite
width. The introduction of a magnetic eld component parallel to the 2DES
will inuence the ground subband wave function
0
(z) of the conned 2DES.
Tilting the sample in a magnetic eld, squeezes the wave function in the
z-direction, thereby making the 2DES more two-dimensional. Variational
calculations suggest that, for samples studied here,
0
(z) will almost halve
its z-extension in a parallel eld component of 20 T [49].
The situation is, in fact, more complex then it appears at rst sight.
For the particular situation of a magnetic eld perpendicular to the 2DES,
the quantum mechanical problem for an electron was separated into out-
of-plane and in-plane motions and simple expressions for the eigenenergies
were obtained. For tilt angles ,= 0, the motions in the connement plane
and perpendicular to it are coupled. Additionally, assuming an arbitrary 1D
connement potential, the model is not analytically soluble. The resulting
eigenenergies are labelled by one quantum number plus the spin variable (j
and ), instead of r, N, and . Theoretically, the single-electron problem
in tilted magnetic elds was investigated by several authors [18]. Halonen
et al. [47] considered the inuence of subband-LL coupling on the fractional
QHE excitation gaps.
In the following, we briey remark on how the LL separation depends
with the tilt angle in samples M242 and M280. Elaborate energy level
numerical calculations were performed by Jungwirth and MacDonald. These
authors went beyond the Hartree approximation by including the exchange
and correlation eects within the local density approximation scheme and
take into account the possibility of an in-plane magnetic eld.
40 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.14: Numerical calculation for the Landau level (LL) crossing be-
tween the lowest spin-down LL and the rst spin-up LL in tilted magnetic
elds [sample M242 (dashed curve) and sample M280 (dash-dotted curve)].
[Courtesy of T. Jungwirth]
Referring to Fig. (1.6), we are interested in the crossing between the second
LL for majority spins (j = 1, = 1) and the rst LL for minority spins
(j = 0, = 1). The energy separation between these two levels is
01
and its evolution with the tilt angle is shown in Fig. (1.14) for = 1. The
theory predicts that the tilt angle where this LL crossing occurs is slightly
higher than = 70

for both M242 and M280 samples. This value is weakly


dependent on the form of the exchange correlation potential and is rather
insensitive for 0.8 1. Due to this possible LL crossing, great care has
to be exercised when analyzing tilted-magnetic eld experiments at high .
1.6. SUPPLEMENTAL ASPECTS OF THE QHE 41
Figure 1.15: Longitudinal resistance R
L
vs lling factor ( 2) at T =
50 mK and indicated tilt angles (sample M242). Curves are oset for clarity.
42 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.16: Longitudinal resistance R
L
vs lling factor ( 2) at T =
50 mK and indicated tilt angles (sample M280). Curves are oset for clarity.
1.7. MAGNETOTRANSPORT: RESULTS AND DISCUSSION 43
1.7 Magnetotransport: Results and discussion
1.7.1 Measurements in tilted magnetic elds
The following magneto-transport measurements describe the evolution of the
QHE with the tilt angle for 2 in our samples. Figures (1.15) and (1.16)
show magnetoresistance traces taken at T = 50 mK and various tilt angles
in samples M242 and M280, respectively. A rather complex structure of
the fractional QHE for 2 has been observed in previous tilted magnetic
eld studies on single heterojunctions [21, 31, 32]. In the present data,
several fractional QHE states are readily detectable: = 2/3, 4/3, and
5/3 in sample M280 [Fig. (1.16)], and = 2/3 and 4/3 in sample M242
[Fig. (1.15)]. Only very small changes occurred by tilting the samples around
= 5/3. The R
L
minima at = 2/3 and = 1 became deeper in the case
of tilted samples. With increasing the tilt angle we observe the development
of a peak in the longitudinal resistance at 0.8 instead of the weak
minimum displayed at = 0

. We believe that the weak structure in R


L
for
2/3 1 is simply consumed by the widening of the adjacent QHE
states. This behavior is consistent with the standard model of disorder-
driven QHE transitions [58]. On the other hand, for unpolarized QHE states
(such as = 4/3), the increased total magnetic eld is the main factor, of
more signicance than the eect of B
|
on the z-extent of the electron wave
function. This is clearly reected in our measurements which reveal that the
= 4/3 QHE state weakens with increasing the tilt angle and it is absent
for 65

. In sample M280 we found, in fact, a re-entrant behavior for


the = 4/3 QHE state, as it re-emerges for 72

[21]. The systematic


dierence observed in the behavior of = 2/3 and 4/3 QHE states, suggests
the destruction of the electron-hole symmetry in the fractional QHE in tilted
magnetic elds. The disparity between = 4/3 and = 5/3 states, believed
to be identical if they occur at the same perpendicular magnetic eld in the
(fully polarized) hierarchical model for the fractional QHE, is also evident.
1.7.2 QHE excitation gap measurements
We address here another important aspect of the QHE, namely, how the
measured transport QHE excitation gaps compare quantitatively with the
theoretical predictions. The investigation of integer an fractional QHE exci-
tation gaps has been one of the major objectives in the early studies of the
QHE. The angular dependence of the excitation gaps has been intensively
used to unravel the spin congurations of various fractional QHE states
as a function of the Zeeman energy. As the Zeeman energy is increased,
44 CHAPTER 1. THE QUANTUM HALL EFFECT
spin-unpolarized fractional QHE states undergo transitions ,
where , , and signify zero, partial, and full polarization. The = 4/3
fractional QHE state provided the rst indications for a spin-unpolarized
QHE ground state. Compelling evidence came from experiments of Clark et
al. [21], which mapped out the destruction of the = 4/3 fractional QHE,
followed by its re-emergence upon increasing the tilt angle.
The values of the QHE excitation gaps are usually determined from the
T-dependence of the longitudinal resistance at QHE minima
23
. The frame-
work which is useful in analyzing and interpreting the activation energy data
has been excellently summarized by Boebinger et al. [13]. To extract the
QHE excitation gap

(expressed in K) at lling factor we use the ex-


pression R
L
(T) exp (

/2T) to produce Arrhenius plots [ln R


L
vs 1/T].
Then, we perform linear ts to the midsection of these plots [Figs. (1.17)
and (1.18)]. Physically, this procedure is equivalent to assume that the con-
duction mechanism is by thermal activation of charged quasiparticles across
the QHE excitation gap
24
. Arrhenius plots of data always develop a down-
ward curvature at high T, as a consequence of a high density of excitations.
Additionally, the data can show an upward curvature at low T, so that the
full plot is S shaped
25
. The upward curvature is attributed to hopping
conduction via localized states near the Fermi level. Meaningful ts using
the sum of two conduction mechanisms are not possible due to the reduced
dynamic range of the data. Figure (1.18) shows results from an experiment
on sample M242 in which the T-dependence of the longitudinal resistance
was determined at = 4/3. This was accomplished in two equivalent ways
- either by systematic isothermal magnetic eld sweeps to pick o the R
L
minimum, or by xing the magnetic eld at the minimum, and changing the
temperature. Our raw data was mainly collected by sweeping the mixing
chamber temperature. In the present experiments, the R
L
(T) dependencies
are measured over a range of temperature within the practical limit of our
dilution refrigerator: 40 mK T 2 K. We note that deviations in R
L
from a simple activated behavior mainly originate from the following facts:
(1) the weakly developed minimum resides on top of a weakly T-dependent
background, (2) the xed B diers from the exact value corresponding to
23
We assume that the nuclear magnetization of GaAs QWs does not aect the QHE
excitation gaps through the hyperne interaction (see Chapter 4).
24
The longitudinal conductivity
L
[and, consequently
L
; due to
L
=
L
/(
2
L
+

2
H
)
L
/
2
H

L
] is expected to be proportional to the number of thermally activated,
electrically charged quasiparticles.
25
At low T, for samples mounted in vacuo and in good thermal contact to the
3
He/
4
He
mixture this upward curvature is conspicuously lacking.
1.7. THE EXCITATION GAP 45
Figure 1.17: Arrhenius plots of the longitudinal resistance R
L
for QHE
states at (a) = 1 and (b) = 1/3 (squares) and = 2/3 (solid curve).
Dashed lines represent ts to the activated regions of the data, from which
the excitation gaps

were determined.
46 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.18: (a) R
L
vs B at various xed temperatures. (b) Arrhenius plots
of R
L
at = 4/3 [discrete T (circles) and continuous T (solid curve)].
1.7. THE EXCITATION GAP 47
Table 1.5: The measured transport gaps

for prominent QHE states at


various tilt angles are given in K (sample M280). Theoretical estimates
t

[Eq. (1.19) with C


3
= 0.1 and C
5
= 0.03] are also included.
Landau level lling factor
1/3 2/5 2/3 1 4/3

( = 0

) 4.8 1.2 1.1 20 0.4

( = 30

) 4.6 - 1.6 22 0.4

( = 45

) - - 2.2 26 0.3

( = 60

) - - 2.9 30 -

16 4.8 12 124 8
the longitudinal resistivity minima
26
, (3) errors in the temperature reading
induced either by too fast warming up of the mixing chamber either by
too strong level of samples excitation. The uncertainty in the measured
transport gaps is 10% at low but increases to 30% at high . In
Table (1.5) we have listed the measured excitation gap of several QHE states
at various tilt angles for sample M280. The most consequential among
the investigated QHE excitation gaps is that at = 1. During the late
1980s, QHE excitation gap studies at = 1 clearly established that the
Coulomb energy plays a dominant role, leading to a substantially larger
QHE excitation gap than the single-particle Zeeman splitting [83, 104]. This
aspect, of relevance to the nature of the QHE excitation gap at = 1, will
be addressed separately, in the next chapter. The rst topic to be addressed
here is the evolution of the QHE excitation gaps at the thirds with . Data
collected at = 2/3 shows that the QHE excitation gap rises with increasing
Zeeman energy, whereas at = 1/3 the QHE excitation gap stays constant
upon increasing the Zeeman energy. This dierent angular dependence for
= 1/3 and 2/3 is explained by the theory of subband-LL coupling for the
fractional QHE in tilted magnetic elds [47]. A dissimilar evolution with the
tilt angle is equally observed for = 4/3 and 5/3 fractions. We tentatively
conclude that = 1/3, 2/3, and 5/3 states preserve the same conguration
for all tilt angles, whereas the conguration for = 4/3 at low is .
26
Hysteresis eects in the superconducting magnet cause 0.02 T unrepeatability in the
magnetic eld values.
48 CHAPTER 1. THE QUANTUM HALL EFFECT
Figure 1.19: (a) R
L
vs lling factor (1 2) at T = 50 mK and selected
tilt angles (sample M280). Panel (b) shows the B-dependence of
4/3
. The
grey region indicates the B-range where the = 4/3 fractional QHE is
absent. The corresponding tilt angles are given on the top axis.
1.7. THE EXCITATION GAP 49
The reentrant behavior of the = 4/3 state is displayed in Fig. (1.19). De-
struction of the 4/3 structure for 65

is consistent with as increasing


the Zeeman energy forces the system into a dierent spin state (presumably,
). The 4/3 state collapses from
4/3
= 440 mK to zero over the magnetic
eld range 2.8 6.5 T, and it is absent for 6.5 B 8.8 T. Our observa-
tions are too scant to speculate on the detailed nature of the reentrant 4/3
state
27
.
We now contrast the measured QHE excitation gaps at the thirds with
the theoretical predictions. The observed
3
is smaller by a factor of 4
than the unbound quasiparticle-quasihole pair-creation energy gap
3

t
. It
is generally believed that this discrepancy stems from the combined impact
of three important factors: the nite thickness of the 2DES, the LL mixing,
and the ubiquitous presence of disorder. Since no calculation currently treats
these three corrections simultaneously, we must consider them in cumula-
tive succession. Existing calculations suggest that upon adequate inclusion
of the nite thickness and LL mixing corrections, the QHE excitations gaps
will be reduced by 50%. These corrected theoretical gaps overestimate our
experimental gaps by a factor of 2. This large oset between theoretical and
experimental values is rather disturbing and incorporation of disorder into
theory remains a challenging task. Under the empirical assumption that dis-
order leads to a magnetic eld-independent gap reduction, all theoretically
values need to be shifted down rigidly. Even though a rigid displacement

of about 3 K is able to simultaneously describe the measured QHE exci-


tations gaps at 1/3, 2/3, and 4/3, it should be noted that this simple model
is questionable when

, for which thermal excitations do not yield


a simple exponential behavior. Finally, we remark on the measured QHE
excitation gaps at = 2/5 and 3/5. Experimentally, we nd at = 0

that

2/5
/
1/3

3/5
/
2/3
0.25, compared with an expected ratio
5
/
3

of 0.3 from simple theoretical considerations. In spite of the quantitative


inconsistency with the theoretical predictions, it is a pivotal aspect that
the measured QHE excitation gaps compare well with those observed in
conventional heterointerfaces. The realization of high-quality multiple-QW
heterostructures is a magnicent development which made possible the birth
of quantum Hall eect Skyrmions.
27
There are indications that the scenario proposed by Clark et al. [21] directly applies
to our measurements. To complete the QHE picture for 1 < < 2, one may assign the
daughter states 8/5 and 7/5. The 7/5 state with e

= e/5 is absent on increasing B


until the parent 4/3 transition in which e

changes to e/5. The proposed 4/3 state is


(14/15 , 2/5 ). At high B, only the parent 5/3 and the reemergent 4/3 state remain.
50 CHAPTER 1. THE QUANTUM HALL EFFECT
1.8 Summary
In this chapter we have introduced the fabrication and physics of two-
dimensional electron systems, along with some properties of electron-doped
GaAs/Al
x
Ga
1x
As multiple-quantum-well heterostructures, interesting in
their own right.
We have detailed several basic magneto-transport experiments revealing
the quantization of the Hall eect. At certain integer ( = i; i = 1, 2, . . .)
and fractional ( = p/q; p = 1, 2, . . . ; q = 3, 5, . . .) Landau level lling
factors plateaus develop in the Hall resistance together with minima in the
longitudinal resistance. The value of the Hall resistance at the plateaus
h/(e
2
) depends only on fundamental constants of the nature. I gave a
succinct overview of the rich phenomenology of the quantum Hall eects
(QHEs) and briey traced the history of the development of our theoretical
understanding of these remarkable low-temperature phenomena.
Many properties of the QHE states have been probed by magneto-
transport measurements in recent years, and collectively, they make a cogent
evidence for the electron-electron interaction playing an important role at
odd Landau level lling factors (e.g., = 1). The results obtained in tilted
magnetic elds in the fractional QHE regime add a new twist and are con-
sistent with the observation of unpolarized (or partially polarized) fractional
QHE states. It is evident from our measurements that the spin polarization
of various fractional QHE states (e.g., = 4/3) varies with the Zeeman
energy, which favors fully polarized ground states.
The discussion initiated here will be continued in the next three chapters,
which will address more particularly the physics of two-dimensional electron
systems around Landau level lling factor one.
Chapter 2
Quantum Hall Eect
Skyrmions
Piara ziua n care m-am nascut si noaptea care a zis: un prunc
de parte barbateasca s-a zamislit! Ziua aceea sa se faca ntuneric
si Domnului din cer sa nu-i pese de ea si lumina sa n-o mai lu-
mineze. . . Sa se ntunece stelele revarsatului zorilor ei; sa astepte
lumina si nimic sa nu vina si sa nu mai vada genele aurorei. . .
Cartea lui Iov
2.1 General formulation
At Landau level lling factor = 1, the two-dimensional electron system
(2DES) condenses into a fully ferromagnetic aligned quantum Hall eect
(QHE) ground state: all electron magnetic moments line up to point in the
magnetic eld direction. Since spins are bound to mobile charges, the 2DES
could be viewed as an itinerant ferromagnet with a quantized Hall resistiv-
ity: a QHE ferromagnet. What distinguishes the QHE ferromagnet from
other conventional ferromagnets is the physical nature of its low-lying ex-
cited states. If the QHE ferromagnet is excited energetically, disturbing the
electron spins away from their mutually-parallel arrangement, a Skyrmion
(charged spin-texture) results.
51
52 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Skyrmions are not the only type of elementary excitations of the = 1 QHE
ferromagnet. We have seen in the precedent chapter that various elementary
excitations related to charge degrees of freedom exist at = 1. This chapter
will be concerned with the nature of elementary excitations associated to
spin degrees of freedom: What is the character of spin excitations (collective,
topological,. . . )? Are they neutral or charged? Are they separated from the
ground state by an energy gap or not? The answer to these questions is
bewildering as it matters if the system is precisely at = 1 or near = 1.
A brief survey of some central ideas of the theory of ferromagnetism
is a necessary prerequisite for inquiring about spin excitations in the QHE
regime and, in particular, QHE Skyrmions. The reason is that the theory
of ferromagnetism embraces exotic structures spawned by the alliance of
symmetry and dimensionality. Additionally, these magnetic structures can
be strongly aected by topological constraints that arise from the global
symmetry of magnetic moments. If we want to understand QHE Skyrmions,
we must now scrutinize the part played in the theory of ferromagnetism by
ideas such as Symmetry, Dimensionality, Topology, . . .
2.1.1 Symmetry and ferromagnetism
Of all distinct forms of symmetry, the broken symmetry has, perhaps, the
most profound implications in QHE physics. Here is the preface that Philip
W. Anderson, Nobel laureat who fabricated the concept of broken symmetry,
placed before the description of the discovery: Our interest is often focused
on the set of low-energy excited states of a system as the physically most
fundamental property of it. Relationships among symmetry and elementary
excitation spectra, which can be lumped together under the name of theory
of broken symmetry, play a vital role in the theory of almost all forms of
quantum condensation
1
.
The relationship of broken symmetry to elementary excitation spectrum
is the essence of Goldstones theorem
2
: Any system with a broken continuous
symmetry has Goldstone modes. These gapless collective modes are space-
time dependent oscillations in the order parameter eld whose non-vanishing
ground state expectation value denes the broken symmetry. This theorem is
very general and its proof is rigorous. For a useful illustration of Goldstones
theorem, let us take a matter about which none of us feel the slightest doubt:
the isotropic ferromagnet. At zero magnetic eld, isotropic ferromagnets are
1
P.W. Anderson, Concepts in Solids, (World Scientic, Singapore, 1997), p. 175-182.
2
J. Goldstone, Field Theories with Superconductor Solutions, Nuovo Cimento 19,
154 (1961).
2.1. GENERAL FORMULATION 53
described by the rotationally invariant Hamiltonian
H
isotropic
=

i,=j=1
s
i
s
j
, (2.1)
where > 0 is the ferromagnetic exchange integral, s
i
denotes the spin
operator of ith particle, the interaction between every pair of particles is
counted once, which accounts for the factor
1
2
, and the double sum runs
over the indices i and j separately, excluding the value i = j. Above its
ordering temperature T, the system has SO(3) symmetry and no permanent
magnetic moment. But when cooled below its ordering temperature, it
becomes spontaneously magnetized in a particular direction, determined by
random spin uctuations [Fig. (2.1)(a)]. The fully ferromagnetic ground
state has all spins aligned along the same particular direction and is clearly
not rotationally invariant: instead, it possesses SO(2) symmetry. In other
words, the ground state spontaneously breaks the continuous rotational
symmetry. Indeed, there are innitely many alternative ground states, in
which all the spins are lined up together in dierent directions [Fig. (2.1)(b)].
Figure 2.1: (a) Ferromagnetic ground state; below the ordering temperature
all spins are aligned. (b) An alternative possible ground state. (c) Long
wavelength excitation, involving rotation of neighboring spins.
54 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
This situation is accompanied by the presence of gapless collective modes
by virtue of Goldstones theorem. As the total spin operator has a non-
vanishing expectation value in the ground state, we can easily identify the
order parameter as being the local magnetization (alias the local spin den-
sity). Thus, Goldstone modes correspond to space-time dependent oscil-
lations in the local magnetization eld. More precisely, Goldstone modes
are spin-wave excitations for which the energy approaches zero in the long
wavelength limit ( ). Figure (2.1)(c) shows a spin wave of long wave-
length and gives us a further bit of useful physical insight. Over regions of
size a one has approximately a ground state with magnetization in a
particular direction. But corresponding to the existence of the spin wave
excitation, the spins will rotate from region to region, with a characteristic
spatial period of order . Now it requires very little energy to bring about
this excitation provided the spin-spin forces are of nite range. In this case
one can always nd a large enough such that the energy required to excite
the mode is essentially zero. The Goldstones theorem, so transparent in
this case, has profound consequences for QHE systems. In short, a broken
continuous symmetry (rotational, translational,. . . ) signals the presence of
gapless elementary excitations.
2.1.2 Dimensionality and spin waves
Ferromagnetism in electron systems occurs not because of direct dipolar
interaction between electron magnetic moments, but rather derives from a
combination of electrostatic forces, electrons kinetic energy, spin, and statis-
tics. It was Heisenberg who rst realized that electron ferromagnetism is
intrinsically a quantum many-body eect, and proposed the scenario that
spin-independent Coulomb interaction and Pauli exclusion principle gener-
ate direct exchange interaction between electronic spins. Soon after, the
phenomenon of exchange coupling among itinerant electrons was suggested
as having importance for magnetism by Bloch. The collective electron treat-
ment of ferromagnetism enables, for example, an immediate interpretation
for the non-integral values of the atomic moments of ferromagnetic elemen-
tal metals such as iron. In ferromagnetic elemental metals where a narrow
d band is overlapping with a much larger s band, the ferromagnetism may
be attributed to the electrons in the partially lled band corresponding to
the d electron states. Nevertheless, the derivation of ferromagnetism from a
reasonable model of itinerant electrons is a challenging problem in theoreti-
cal physics. A unique possibility is oered by the QHE ferromagnet at = 1
to understand electron itinerant ferromagnetism without the complications
2.1. GENERAL FORMULATION 55
imposed by the band theory. The simple reason is that the lowest Landau
level behaves like a band of zero width [53].
The fact that the lowest energy band in the single-electron spectrum is
dispersionless, is not the only dierence between the ferromagnetism of the
2DES at = 1 and, lets say, the ferromagnetism of a piece of bulk iron. It
is well known that the low-lying excited states of conventional ferromagnets
have spin-wave character and determine their thermodynamic properties at
low T. For example, the spontaneous magnetization at zero magnetic eld
depends on the form of the band, the magnitude of the exchange inter-
action, and must be calculated on the basis of Fermi-Dirac statistics. In
the spin-wave approximation, the magnetization approaches the saturation
as T
3/2
(the so-called T
3/2
-Bloch law)
3
. For spin-wave phenomena, both
reduced dimensionality and nite magnetic elds become paramount. For
instance, Mermin and Wagner rigorously demonstrated that there can be
no spontaneous magnetization at any nite T for isotropic two-dimensional
(2D) short-range spin models
4
. With respect to spin-wave phenomena, fer-
romagnetism in 2DESs at = 1 deviates radically from the ferromagnetism
of bulk iron: in the former case, electrons experience a nite magnetic eld
B and their dynamics is two-dimensional.
In order to make the concept of spin waves in QHE ferromagnets more
tangible, it is useful to elaborate rst on standard 2D spin-wave excita-
tions. If
H
> 0 is the ferromagnetic exchange integral, s
i
denotes the spin
operator at the site R
i
on the 2D lattice, and the Zeeman interaction is taken
into account, then the 2D insulating isotropic (Heisenberg) ferromagnet is
described by the Hamiltonian
H
H
=

H
2
A

i,=j=1
s
i
s
j
E
Z
A

i=1
s
z
i
. (2.2)
Here we consider that the magnitude of spin is s, B lies along z-axis, E
Z
is
the Zeeman splitting, and the total number of spins is ^. The ground state
3
The ferromagnetic ideal survived the discovery of amorphous metals in 1960s, because
ferromagnetic exchange coupling does not depend in any fundamental way on the existence
of a crystalline lattice. Details may change in the amorphous state: the T-dependence of
the spontaneous magnetization is modied owing to a distribution of exchange interac-
tions; yet the critical behavior is not fundamentally altered. We also note that disorder
in QHE systems plays a role similar to the band dispersion in the band theory of ferro-
magnetism.
4
N.D. Mermin and H. Wagner, Absence of Ferromagnetism and Antiferromagnetism
in One- or Two-Dimensional Isotropic Heisenberg Models, Phys. Rev. Lett. 17, 1133
(1966).
56 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
is the one with all spins aligned along z-axis and parallel to each other
H
H
[
1

2
. . .
A
) = E
0
H
[
1

2
. . .
A
), (2.3)
and its energy E
0
H
is
E
0
H
=

H
2
A

i,=j=1
s
2
E
Z
A

i=1
s. (2.4)
To construct the low-lying excited states one starts by examining a state
[R
j
) diering from the ground state only in that the spin at site R
j
has its
component along the B-axis reduced from s to s1. This state is generated
by the so-called local spin-lowering operator s

(R
j
)
[R
j
) =
1

2s
s

(R
j
)[
1

2
. . .
j
. . .
A
). (2.5)
By making a superposition of these states, we obtain a spin-wave state of
wavevector q (lying inside the Brillouin zone)
[q) S

q
[
1

2
. . .
A
) =
1

^
A

j=1
e
iqR
j
[R
j
)
=
1

^
1

2s
A

j=1
e
iqR
j
s

(R
j
)[
1

2
. . .
A
).
(2.6)
The excitation energy of the state [q), generated from the ground state by
S

q
[the Fourier transform of the local spin-lowering operator s

(R
j
)] is
E
H
(q) = s
H

p
_
1 e
iqR
p
_
+E
Z
. (2.7)
Here

p
represents the sum over the nearest-neighbor lattice vectors. For
a square 2D lattice of constant a, at small magnitude q of the wavevector,
the dispersion of the collective mode is quadratic
E
H
(q) s
H
q
2
a
2
+E
Z
. (2.8)
If we wish to pursue the analysis further, and use Eq. (2.8) to calculate
the T-dependence of the magnetization at B = 0, we shall arrive at the
remarkable result that at any nite T so many spin waves are excited that
the magnetization is completely eliminated. The density of states of a branch
2.1. GENERAL FORMULATION 57
of excitations with energy E
H
(q) scales as q dq and the occupancy of the
low-lying modes diverges logarithmically at q 0. Any B ,= 0 can make
the frequency of the q = 0 spin-wave mode nite. Consequently, the T-
dependence of the magnetization is essentially T ln (k
B
T/E
Z
), where E
Z
acts as a cuto energy. In the above example, the probability of lowered spin
being found at a particular site R
j
in the state [q) is 1/^; i.e., the lowered
spin is distributed with equal probability among all the spins. We shall
see soon that, although the concept of a spin wave in QHE ferromagnets is
very dierent from that of a single reversed spin distributed coherently over
a large number of otherwise aligned spins, the dispersion of the spin-wave
mode at small q is still quadratic.
2.1.3 The abomination of topology
The Kosterlitz-Thouless phase transition
According to the Mermin-Wagner theorem the conventional long range or-
der (positional spin alignment) at nite T for rotationally invariant 2D spin
models is prevented by thermal uctuations. However, the Mermin-Wagner
theorem does not rule out the possibility of some other type of ordering or
nite-T phase transition in 2D systems. During the early 1970s, Kosterlitz
and Thouless showed that a phase transition occurs at a nite tempera-
ture (T
KT
) in 2D XY -ferromagnets with short-range interactions
5
. These
authors considered the Hamiltonian for a system of planar spins with unit
modulus on a square 2D lattice
H
KT
=
KT

ij)
s
i
s
j
=
KT

ij)
cos (
i

j
), (2.9)
where
KT
> 0 is the ferromagnetic exchange integral,

ij)
means that
the sum is taken over all pairs of nearest-neighbor sites, and s
j
is the spin
operator at the site j, represented by s
j
= cos
j
e
x
+sin
j
e
y
, with
j
be-
ing the in-plane orientation angle. The interesting physics is brought about
by the vortices of XY spins. An isolated vortex in the XY -model, shown
in Fig. (2.2), has several remarkable properties: (1) around the center of
the vortex the spin scan rotate along closed contours with an angle 2m,
5
J.M. Kosterlitz and D.J. Thouless, Ordering, Metastability, and Phase Transitions in
Two-Dimensional Systems, J. Phys. C: Solid State Phys. 6, 1181 (1973). At any nite T
the total magnetization is zero. Yet, a clear qualitative change happens within the system
at T = T
KT
, signaled by dierent behaviors of the spin-spin correlation function below
and above T
KT
.
58 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
with m an integer number (the so-called vorticity or strength of the vor-
tex), (2) the total magnetic moment of the system containing such a vortex
structure is zero, (3) the spin ordering clearly diers from random order-
ing, and (4) moving along a vortex contour the angle deviations between
nearest-neighbor spins is small. The energy associated with the variation of
the orientation angle in such a vortex structure can be written as
E
vortex
=

KT
2

ij)
(
i

j
)
2
. (2.10)
For an isolated vortex of radius R, it turns out that
E
vortex

KT
ln (R/a), (2.11)
Figure 2.2: Schematic illustration of an isolated vortex in the XY model.
We can take the center of the vortex to be located on a dual lattice whose
sites lie at the centers of the squares of the original lattice. The minimal
contour around the center of the vortex passes through only four sites. The
strength of the vortex is m = 1.
2.1. GENERAL FORMULATION 59
where a is the vortex core radius (alias the lattice constant). At low T, the
free energy variation per single vortex is
F
vortex
(
KT
2k
B
T) ln (R/a). (2.12)
This quantity depends logarithmically on the size of the system, so that the
concentration of free vortices at low T is zero in the thermodynamic limit.
The critical temperature associated with the creation of a single vortex is
T
KT
[/(2k
B
)]
KT
, as F
vortex
< 0 for T > T
KT
. The immediate question
that comes in mind is what kind of phase exists below T
KT
? A very clever
way to see this is to draw an analogy between the system of vortices and a
2D system of electrical charges, and notice that a pair of opposite charges -
a dipole - has a nite total energy, because at large distances the electrical
elds from each charge almost cancel. Thus, it is presumed that in the low-T
phase, vortices can only appear in bound pairs, with vorticities of dierent
sign (vortex anti-vortex pairs)
6
. As the density of vortex anti-vortex
pairs raises with the temperature, the pairs begin to screen each other and
eventually unbind. In the high-T phase, the system behaves like a plasma
composed of an equal number of free, screened charge of each sign. This
is the essence of the Kosterlitz-Thouless phase transition
7
. This particular
physics - pertaining to phase transitions in 2D systems with (magnetic)
short-range interactions, will appear a little while further, in our discussion
of QHE ferromagnetism near = 1.
The nature of this type of phase transition is topological because the
most prominent feature of the low-T phase is the presence of non-trivial
topological order, rather than the existence of a conventional order param-
eter. As pointed out by Kosterlitz and Thouless, the idea of topological
ordering appeared for the rst time in the dislocation theory of lattice melt-
ing. In this theory, it is assumed that a liquid close to the freezing point
has a local structure similar to that of a solid. In the liquid phase, there is
a concentration of free dislocations, which can move under the inuence of
an arbitrary small shear stress. Below the melting point, there are no free
dislocations; the dislocations are coupled in pairs, so that the dislocations in
the same pair are described by equal and opposite Burgers vectors. There-
fore, the solid-to-liquid phase transition is, in fact, a Kosterlitz-Thouless
phase transition between a state with dislocation pairs and a state with free
topological defects
8
.
6
As a consequence, below T
KT
the total magnetization of the system is indeed zero.
7
An important remark here is that T
KT
decreases when the number of particles per
unit area decreases.
8
The theory of lattice melting is much easier to apply in two dimensions than in tree
60 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Classical treatment of a Skyrmion
Before sketching the topological nature of the Skyrmion, it is the breadth
generality of the classical theories for the ferromagnetic state that I wish
to stress. Let me say immediately that one of the main reasons for the
continued interest in ferromagnetism is a rare opportunity of simultane-
ously describing complex phenomena occurring in the same system both in
classical and quantum-mechanical terms. The classical approach takes the
quantum mechanical concept of a spin and treats it as if it were a classical
vector. While the quantum theory proceeds from the formulation of the
spin Hamiltonian, the principal assumption of the classical theory is that
the ferromagnetic state is unambiguously describable
9
by the magnetiza-
tion vector M. In general, the magnetization as a function of space and
time M(r, t) is a solution of the phenomenological equation
dM
dt
= MB
eff
M
dM
dt
. (2.13)
The precession and the motional damping of the vector M are described by
the gyromagnetic factor and the relaxation constant , respectively [59].
The eective magnetic eld in Eq. (2.13) is the variational derivative of the
energy functional with respect to the vector M
B
eff
=
E[M]
M
. (2.14)
This equation has the integral of the motion M
2
= constant, consistent with
the assumption that the length of the vector M in a ferromagnet represents
its equilibrium characteristic. In the ground state the value of M is the
spontaneous total magnetization M
0
. The focus here is on the dynamics
of the magnetization when the time scales involved are such that M main-
tains its ground state magnitude during the motion. Equation (2.13) may
have special solutions which are generated by its mathematical structure
and related to the global symmetry of M(r). Such solutions are usually
topologically specic ones and they are called topological solitons [59]. In
since a dislocation is associated with a point rather than a line.
9
We shall assume that the total magnetization of the thermodynamic system M(r; B)
is single valued for any r and B. Systems for which M(r; B) is not single valued are
said to exhibit hysteresis; most ferromagnetic systems have this property. Hysteresis is
associated with a magnetic heterogeneity of the sample, the separate regions being known
as domains. The analysis we shall develop is generally applicable within a single domain.
This assumption is valid for the = 1 QHE ground state as long as it behaves like a
homogeneous and isotropic quantum liquid.
2.1. GENERAL FORMULATION 61
order to understand the situation with topological solitons, a unit vector
m(r) = M(r)/M
0
is introduced which maps the ferromagnet space onto the
surface of a unit sphere.
Lets specialize now to the case of isotropic 2D ferromagnets. In terms
of the unit vector eld m(r), the energy functional is given by
E[m(r)] =

s
2
_

a
m

(r)
a
m

(r) d
2
r, (2.15)
where
s
is the ferromagnetic spin stiness which has units of energy, Greek
indices refer to the directions in the spin space ( = x, y, z), and Latin
indices refer to the directions in the coordinate space. The energy functional
E[m(r)] (the so-called minimal gradient energy) express the cost due to
the loss of Coulomb exchange when the spin orientation varies with the
position and it is the rst term in the Taylor expansion of dipolar magnetic
interaction which preserves the rotational symmetry. In close analogy to
Skyrmes theory
10
, the dynamics of the isotropic ferromagnet could be
studied within the Lagrangian formalism, using the scalar elds m

(r) with
the constraint m(r) m(r) = 1. The static congurations of an innite
continuum system can be obtained by solving Euler-Lagrange equations that
follow from the Lagrangian
L =

s
2
_

a
m

(r)
a
m

(r) d
2
r +
L
_
[m

(r)m

(r) 1)] d
2
r, (2.16)
where
L
is a Lagrange multiplier that enforces the length constraint. The
unit magnetization vector eld obeys a non-linear equation
11

2
m(r)
_
m(r)
2
m(r)

m(r) = 0. (2.17)
Notice that the constraint introduced a completely new twist to the problem.
Without the constraint, Eq. (2.17) reduces to the much simpler Laplace
10
In particle physics, the goal of the eld theory is to explain the rich diversity of
experimental phenomena with a very small number of fundamental elds that propagate
under the inuence of a highly symmetric Lagrangian. The Skyrmion was born with a
famous conjecture by Skyrme - that a particular class of non-linear eld theories is relevant
to the description of elementary particles such as mesons. The mesonic eld obeys a non-
linear dierential equation which has solutions that approximate to travelling waves. In
addition, the equation has a static solution which could be interpreted as a particle
(Skyrmion). [T.H.R. Skyrme, A Non-Linear Field Theory, Proc. Roy. Soc. (London) Ser.
A 260, 127 (1961).]
11
This equation is, of course, identical to the one deduced from Eqs. (2.13 - 2.15) within
the assumption that = 0 and
dM
dt
= 0.
62 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
equation. The standard Skyrmion solution of Eq. (2.17) is
m
x
(r) =
2x

2
+r
2
=
2r cos ( )

2
+r
2
m
y
(r) =
2y

2
+r
2
=
2r sin ( )

2
+r
2
m
z
(r) =
r
2

2
+r
2
, (2.18)
where (r, ) are the polar coordinates in the coordinate plane, is a con-
stant that controls the in-plane spin orientation, and is the size of the
Skyrmion [95]. This is illustrated in Fig. (2.3), in which spins have been
joined to sites of a square lattice. The magnetization varies smoothly from
down (at the center) to up (at innity) as function of radius and angle with
m(r) lying in the xy plane at a distance from the center of the Skyr-
mion. The energy corresponding to this solution is E
1
Sk
= 4
s
and does
not depend on parameters and . In few words, borrowed from the math-
ematical language, the standard Skyrmion is a eld conguration m(r) that
wraps around the O(3) order parameter sphere exactly once and thus, it is
described by a spatially extended topological charge that integrates to unity.
This can be more ample enunciated, using concepts of ordinary topology, as
follows. A quantity called topological charge density is dened as

top
(r) =
1
8

uv
m(r)
_

u
m(r)
v
m(r)

. (2.19)
Since the Jacobian converting area in the plane into the solid angle on the
sphere is 4 times the topological charge density,
top
(r) plays a key role
for the wrapping of the plane onto a sphere. Associated to the topological
charge density is the topological charge given by
Q
top
=
1
8
_

uv
m(r)
_

u
m(r)
v
m(r)

d
2
r. (2.20)
The topological charge must be always an integer since it counts the number
of times the compactied plane is wrapped around the O(3) order parameter
sphere by the mapping m(r). The Skyrmion conguration of Eq. (2.18)
wraps the plane around the O(3) order parameter sphere exactly once and
henceforth, has Q
top
= 1. The importance of the topological charge for the
QHE will become clear shortly, but one aspect is already worth noting at this
stage of discussion
12
. The topological charge is a topological invariant, i.e.,
12
The corresponding anti-Skyrmion has Q
top
= 1. It is of the same form but with
m
y
(r) m
y
(r) and
top
(r)
top
(r). Skyrmions with higher topological charge may
exist. Their energy is E
Q
top
Sk
= 4
s
[Q
top
[ [95].
2.1. GENERAL FORMULATION 63
Figure 2.3: Vector representation of a Skyrmion spin texture on a square
lattice. Unlike a XY vortex there is no singularity at the origin because the
spins are able to rotate out of the plane. The spin is down at the center and
gradually turns up at innite radius. [Courtesy of N.R. Cooper]
64 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
it remains stable against smooth continuous distortions of the eld m(r).
For example, spin waves could travel through the Skyrmion and Q
top
does
not change, i.e., the energy of the Skyrmion remains the same.
The topological phase
Before we go on, it is necessary for me to be relieved of a certain anxiety.
This diculty is brought about by the quantum treatment of spin dynamics
and can be elucidated by the following argument [39, 97]. Lets take the
Lagrangian formalism as a convenient formulation of the laws of dynamics
and look at a spin s = sm precessing under the inuence of a magnetic eld
B. Here s is the spin length and the unit vector m denotes the direction of
spin. The aforementioned diculty arises from the fact that the description
of the system will be given in terms of the time-dependent history of the unit
vector m.
The Lagrangian that reproduces the precession of the quantum spin
should contain (1) the Zeeman interaction term sm

and (2) a term


that enforces the length constraint
L
(m

1). Here is the gyro-


magnetic factor, refers to the directions in the spin space, and
L
is the
Lagrange multiplier. When the orientation of a spin is moved adiabatically
around a closed loop, the system picks up a Berrys phase
13
. Henceforth, a
new term should be added to the Lagragian to account for this quantum ef-
fect. Dening in the spin space a velocity vector s m and a vector potential
A which satises m =
m
A, the Berrys phase term in the Lagrangian
can be expressed as
14
s m

. (2.21)
To sum up, the Lagrangian of the precessing spin turns out to be
L
precession
= s m

+sm

+
L
(m

1). (2.22)
13
The adiabatic principle of Ehrenfest stipulates that any system adiabatically trans-
ported around a closed circuit ( in the parameter space will return in its original state
(of energy E). Its internal clocks will register the passage of time t; this is the origin of
the dynamical phase factor: exp (iEt/). Additionally, the system records its history
in a geometrical way, expressed by a geometrical phase factor: exp [i(()]. [M.V. Berry,
Quantal Phase Factors Accompanying Adiabatic Changes, Proc. Roy. Soc. (London) Ser.
A 392, 45 (1984).] The Berrys phase writes as (() = 2
R
C

top
(r) d
2
r.
14
There is a close analogy between the Berrys term in the quantum Lagrangian of a
precessing spin and the Lagrangian of a charged particle moving under the inuence of a
real magnetic vector potential A. If the particle is massless, its velocity is v, and we set
the electrical charge to 1, then the Lagrangian reduces to the rst order time derivative
term v A. Thus, the Lorentz force is disguised in the Magnus force, proportional to
the vector product of the spin velocity and the local spin magnetization [39, 97].
2.2. QHE FERROMAGNETISM AT = 1 65
This Lagrangian correctly reproduces both the Berrys phase and the equa-
tion of motion for the quantum spin whose Hamiltonian is given by
H
precession
= sm

. (2.23)
We shall see soon how this weird physics ts into the theory of QHE ferro-
magnetism.
2.2 QHE ferromagnetism at = 1
2.2.1 Spin-wave excitations
The spin-wave mode of the QHE ferromagnet at = 1 could be derived
without great mathematical eorts if one uses the convenient formulation
of quantum mechanics within the subspace of the lowest Landau level
15
.
In this formalism, the Hamiltonian for ^ interacting 2D electrons can be
expressed solely in the terms of the projected density operator
k
as
H =
1
2

k
_
V (k)
k

k
ne

1
2
[k[
2
l
2
B

, (2.24)
where V (k) is the Fourier transform of the electron-electron interaction po-
tential V (k) =
_
V
ee
([r r
t
[)e
ik[rr

[
d
2
r. An explicit microscopic wave
function at = 1, which factorizes into the Laughlin spatial wave function
for the lled spin-polarized Landau level
L
1
[z] and the fully symmetric spin
15
For a 2DES with area A and A particles, the Fourier transforms of the charge and
total spin density operator, projected to the lowest Landau level, have the form

q
=
1

A
N
X
j=1
e
iqr
j
=
e

1
4
|q|
2
l
2
B

A
N
X
j=1

q
(j),
S

q
=
1

A
N
X
j=1
e
iqr
j
s

(r
j
) =
e

1
4
|q|
2
l
2
B

A
N
X
j=1

q
(j)s

(r
j
).
Here
q
(j) = e
iql
B

z
j

i
2
q

l
B
z
j
is the so-called magnetic translation operator, which
translates the jth particle a distance l
2
B
(e
z
q). The dimensionless complex coordinate
for the jth particle is z
j
, l
B
=
p
/(eB) is the magnetic length, and refers to directions
in the spin space. Note that, because of the itinerant character of the electrons, the
magnitude of the wavevector q could be arbitrary big. The projected spin and charge
density operators do not commute. This implies that within the lowest Landau level, the
dynamics of spin and charge are entangled [39].
66 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
wave function [
s
), was given in the previous chapter
[z] =
A

i<j
(z
i
z
j
)
A

k=1
e

1
4
[z
k
[
2
[
1

2
. . .
A
). (2.25)
In the standard procedure [see Eq. (2.6)], low-lying excited states, labeled
by the wavevector q, are obtained from the many-body ground state [z]
as

q
[z] S

q
[z] =
A

j=1
e
iqr
j
s

(r
j
)[z], (2.26)
where S

q
is the Fourier transform of the local spin-lowering operator s

(r
j
)
and overline indicates projection onto the lowest Landau level. The energy
of the spin-wave mode is given by
E
sw
(q) =
1
(2)
2
_
e

1
2
[k[
2
l
2
B
V (k)
_
1 cos [e
z
q k]l
2
B

_
d
2
k. (2.27)
For pure Coulomb interactions [V
ee
([r r
t
[) = e
2
/([r r
t
[)] and in the
presence of Zeeman splitting (E
Z
= [g

[
B
B), the dispersion of the spin-
wave mode at = 1 has the following compact analytical form
E
sw
(q) = [g

[
B
B +
e
2
l
B
_

2
_
1/2
_
1 exp (q
2
l
2
B
/4)I
0
(q
2
l
2
B
/4)

, (2.28)
where I
0
is the modied Bessel function of the rst kind [95]. This is illus-
trated in Fig. (2.4) using the estimates for sample M242 given in Table (1.3).
The asymptotic limits of Eq. (2.28) are
E
sw
(q)
q0
= [g

[
B
B +
e
2
l
B
4
_

2
_
1/2
q
2
, (2.29)
E
sw
(q)
q
= [g

[
B
B +
e
2
l
B
_

2
_
1/2
= E
Z
+E
xc
. (2.30)
For q 0 the spin-wave mode energy has a quadratic q-dependence (similar
to 2D isotropic ferromagnets on a lattice [Eq. (2.8)]). Since E
xc
E
Z
,
E
sw
(q ) is a constant of the order of the Coulomb energy [E
C
=
e
2
/(l
B
)]. The physical picture of a spin ip process is uncanny. At small q,
simultaneously with a ipped spin, a bound quasiparticlequasihole pair is
created. At large q, when a spin is ipped, the quasiparticle is concomitantly
2.2. QHE FERROMAGNETISM AT = 1 67
Figure 2.4: Dispersion relation of the spin-wave mode at = 1 [Eq. (2.28)].
Parameters E
Z
= 1.8 K and E
xc
= (/2)
1/2
E
C
= 153 K correspond to
sample M242 [Table (1.3)]. Surprisingly, in the light of the strong magnetic
elds, the physics associated with the spontaneous magnetization at = 1
is experimentally accessible because E
Z
is weak compared to E
C
, and can
even be tuned to zero [75, 92].
moved far away from the quasihole. Interestingly, the Lagrangian formalism
could be used to obtain the dispersion relation of the spin-wave mode at
= 1. This highlights the eld-theoretical approach as truly valuable for
the discussion of = 1 QHE ground state. We have previously seen that
when the spin adiabatically precesses under the inuence of a magnetic eld,
its Lagrangian contains a Berrys phase term. For a 2DES with an electron
density n the Berrys phase term is
L
Berry
= sn
_
m

(r) /

[m(r)] d
2
r, (2.31)
68 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
where describes the directions in the spin space, s =
1
2
is the spin length,
and m(r) is the local magnetization unit vector eld. The Coulomb and
Zeeman terms are
L
Coulomb
=

s
2
_

a
m

(r)
a
m

(r) d
2
r, (2.32)
L
Zeeman
= sn[g

[
B
_
m

(r)B

d
2
r. (2.33)
In the long wavelength limit, provided that m
z
(r) 1 and m(r) m(r) = 1,
the spin-wave Lagrangian L
sw
= L
Berry
+ L
Coulomb
+ L
Zeeman
leads to the
dispersion relation
E
sw
(q) = [g

[
B
B +

s
ns
q
2
. (2.34)
By comparing the expression for the spin-wave mode [Eq. (2.29)] to the
dispersion relation from the above equation, we can identify the 2DES spin
stiness at = 1 as

s
=
1
16

2
e
2
l
B
. (2.35)
Estimates of the spin stiness at = 1 for our measured samples are given
in Table (1.3). Skyrmions may exist in planar conventional ferromagnets.
Ordinary ferromagnets, however, have a large spin stiness; consequently,
Skyrmions are irrelevant for their observable properties.
2.2.2 Topological excitations (QHE Skyrmions)
The way lies now open to develop a Lagrangian formalism for topological
excitations of the = 1 QHE ground state. To nd what a QHE Skyrmion
is, we must include the manifest connection between the spin and charge
at = 1. In a self-consistent description, each electron at = 1 experi-
ences a strong local exchange eld produced by the rest of electrons. As it
propagates under the exchange eld, its spin adiabatically follows the local
orientation of the (slowly spatially varying) spin-textured background. This
adiabatic spin dynamics gives rise to a geometrical phase for closed paths.
The non-trivial spin background in which the electron moves is characterized
by a unit topological charge and produces a Berrys phase which is equal
to the change in the Aharonov-Bohm phase produced by the insertion of
one quantum of ux
0
. For incompressible QHE ground states at lling
factor , Laughlin argued that additional ux into the system creates an
extra charge Q = e/
0
[67]. Consequently, the real charge of a spin
texture with topological charge Q
top
= 1 at = 1 is sharply quantized to the
2.2. QHE FERROMAGNETISM AT = 1 69
elementary electron charge. Spatial spin uctuations may produce a modu-
lation of the topological charge density. Because of the equivalence between
the topological and real charge, we should not lose sight of the eect of
physical charge density uctuations which cost Hartree energy. The Hartree
energy is the next leading term in the spin-texture energy functional given
by Eq. (2.15). Armed with this new knowledge, we can write the eective
Lagrangian that describes the long-wavelength and low-energy physics of
the QHE ferromagnet at = 1 as
L
eff
= L
Berry
+L
Zeeman
+L
Coulomb
+L
Hartree
. (2.36)
If (r) represents the deviation of the charge density from the uniform
background density, the Hartree term in the eective Lagrangian is given by
L
Hartree
=
1
2
_
d
2
r
_
(r)
1
[r r
t
[
(r
t
) d
2
r
t
. (2.37)
Skyrmions are static solutions of L
eff
and can be obtained by energy min-
imization of the dominant (Coulomb) term as described previously. The
minimal gradient energy of the single Skyrmion is E
1
Sk
= 4
s
= E
xc
/4.
Therefore, forgetting for an instant the Zeeman energy, the QHE excitation
gap at = 1 is E
xc
/2. It is two times smaller than the quasiparticle-
quasihole Laughlin pair creation energy [see Sec. (1.5)] and it is nite as
long as the spin stiness is nite.
Quantum Hall eect Skyrmions are quasiparticles characterized by a
nite size, charge, spin, and statistics. Lets peer rst into the size and spin
of Skyrmions. The Zeeman and Hartree terms in the eective Lagrangian
control the stability of individual Skyrmions towards shrinking or growing
in size. The Hartree energy prefers to expand the Skyrmion size, as for large
Skyrmions (r) is small. The Zeeman energy tends to shrink Skyrmions
since it promotes polarization along the magnetic eld direction and tries
to stop the in-plane polarization. The scale invariance of E
1
Sk
is broken by
the presence of Zeeman and Hartree terms, whose competition yields an
optimal size for Skyrmions. The ratio g between the Zeeman splitting and
the Coulomb energy is an important, experimentally accessible parameter,
which controls the size of the Skyrmion.
These qualitative considerations are followed by some relevant quantita-
tive results. First, for a vanishing small g, Sondhi et al. [95] proposed the
70 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
following formulae for the size and energy E

Sk
of the Skyrmion
_
/l
B
_
3

(
l
1
l
B
( g[ ln g[)
1
,
E

Sk
E
1
Sk

e
2
l
B
_
(
l
2
l
B
_
1/3
( g[ ln g[)
1/3
,
(2.38)
where (
l
1
and (
l
2
are positive constants with units of length. One can eas-
ily see that the Skyrmion size diverges in the asymptotic limit of g 0.
Henceforth, the number of ipped spins within a Skyrmion diverges in the
limit of vanishing g. At nite Zeeman coupling, the optimal size ( l
B
)
of a Skyrmion could be obtained by considering the nite-size corrections to
E
1
Sk
in a system with linear dimension R
E

Sk
E
1
Sk
= (
E
1
_

R
_
2
+ (
E
2
_
l
B

_
, (2.39)
where (
E
1
and (
E
2
are positive constants with units of energy [80]. Minimizing
E

Sk
given by Eq. (2.39) with respect to gives R
2/3
l
1/3
B
. Accordingly,
the number of reversed spins scales as
2
ln (R/). Finally, of great im-
portance with regard to the experiments, theory also predicts that above a
critical g value ( g
c
) Skyrmions evolve into single spin-ips. In this limit, the
size of Skyrmions approaches l
B
. For an ideally thin, disorder-free 2DES,
and with no Landau level mixing, calculations [95] give a lower bound of
g
c
= 0.054.
One may ask how relevant are Skyrmions for real, disordered 2DESs in
GaAs-based heterostructures. Because Skyrmions carry physical electrical
charge, unlike the case of a regular ferromagnet, a scalar potential may
induce the formation of Skyrmions in the ground state even at T = 0. The
energy of a Skyrmion in a random disorder potential V
d
(r) is
E

Sk
(r
0
) =
_
(r r
0
, )V
d
(r) d
2
r, (2.40)
where r
0
is the position of the Skyrmion, (r, ) = e
top
(r, ), and

top
(r, ) = [/(r
2
+
2
)]
2
/ is the topological charge density for the stan-
dard Skyrmion of size . The optimal size of the Skyrmion could be obtained
by adding this (no less interesting) term to E

Sk
[Eq. (2.39)] and subsequent
minimization with respect to . Few other comments are in order here. In
the previous chapter we have seen that it is essential for the observation of
the QHE that charged quasiparticles are localized by disorder. In the T 0
2.2. QHE FERROMAGNETISM AT = 1 71
limit, it is believed that Skyrmions are pinned by the disorder and do not
contribute to the charge transport in the = 1 QHE plateau
16
. It is also
expected that Skyrmion anti-Skyrmion pairs will be thermally activated
17
and hence exponentially rare at low T.
We take up now the subject of Skyrmion statistics. The eective La-
grangian [Eq. (2.36)] may also contain the Hopf term [87], reecting that
the ground state does not have time-reversal symmetry and parity invari-
ance. The Hopf term is given by
L
Hopf
=
N
Hopf
32
_

ijk
B
i
F
jk
d
2
r, (2.41)
where
ijk
is the completely antisymmetric tensor, the prefactor N
Hopf
is a
topological invariant, B
i
is an auxiliary gauge eld, and F
ij
=
i
B
j

j
B
i
=
m(r) [
i
m(r)
j
m(r)]. The total spin of the Skyrmion has contributions
from both the Berry term and the Hopf term in the eective Lagrangian. In
the long wavelength limit the contribution due to the former dominates. The
statistics of the Skyrmion, on the other hand, is completely determined by
the prefactor N
Hopf
. This is found to be equal to and the statistical phase
of the Skyrmion is . Consequently, Skyrmions at = 1 obey Fermi-Dirac
statistics
18
.
2.2.3 Further aspects of QHE Skyrmions
While the eld theoretic approach [95] is very successful in qualitatively ex-
plaining the long-wavelength and low-energy physics of QHE ferromagnets,
it fails when it comes to quantitative comparison with experiments. The
16
Even given the fact that the Skyrmions are pinned by the disorder, the question arises
whether the existence of gapless spin waves leads to dissipation and hence destroys the
QHE in the limit of g 0. It is likely that dissipationless ow survives the presence of
spin waves, since they are neutral and hence, irrelevant to charge transport.
17
At nite temperatures, the T-dependence of Skyrmion density may exhibits a non-
Arrhenius behavior due to the presence of disorder.
18
A priori, similar charged spin-texture excitations may occur around other fractional
lling factors with incompressible ferromagnetic many-body ground states. For instance,
at = 1/3, Skyrmions are anyons and obey fractional statistics. However, calculations
of Skyrmion energies near = 1/3 show that Skyrmions are present only for very small
Zeeman couplings, approximately 10 times smaller compared to those at = 1. [R.K.
Kamilla, X.G. Wu, and J.K. Jain, Skyrmions in the Fractional Quantum Hall Eect, Solid
State Commun. 99, 289 (1996).] I should mention that Skyrmions are also expected at
odd integer lling factors such as = 3. Sadly, calculations near = 3 also reveal that
Skyrmions are relevant only for very small g, approximately 10 times smaller compared
to g at = 1.
72 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
failure of the classical eld theory to predict the correct Skyrmion physics
at short length scales originates in the truncation of the gradient expansion
of the energy functional E[m(r)]. On the other hand, Hartree-Fock calcula-
tions [23, 36] are in essence self-consistent mean-eld calculations with the
order of the gradient expansion taken to innity and hence more successful
at predicting the small-size Skyrmion physics.
In a nutshell, the Hartree-Fock formalism used in Skyrmion studies is
as follows. The form of the wave functions is essentially dictated by the
symmetry of the classical Skyrmion solutions which are invariant under the
action of L
z
S
z
for the Skyrmion (anti-Skyrmion). Here L
z
and S
z
denote
the z-components of the total orbital and spin angular momentum, respec-
tively. This property, combined with the lowest Landau level occupancy
requirement, uniquely picks the form of the Hartree-Fock wave function
[
HF
) =

m=0
_
u
m
a
+
m
+v
m
b
+
m+1
_
[0), (2.42)
where [0) is the vacuum state, a
+
m
creates a spin-down electron and b
+
m+1
creates a spin-up electron in the mth angular momentum state. The Hartree-
Fock procedure involves the minimization of the average energy with respect
to the variational parameters u
m
, v
m
, subject to the constraint that the
wave function is normalized. The expectation value of the total spin operator
in the state described by Eq. (2.42) describes a spin-texture with Q
top
= 1
provided that u
m
varies slowly with m from u
m=0
= 1 to u
m
= 0. The
total number of reversed spins
19
in this wave function is K =

m=0
[u
m
[
2
.
The K = 0 Hartree-Fock quasihole is obtained by choosing u
m
= 0 for all m.
The spin of the K = 0 Hartree-Fock quasihole is one half, consistent with
a single spin ip. It represents the change in the total spin of the system
when it is introduced to the fully spin-polarized Landau band. Similarly,
Skyrmions (or anti-Skyrmions) involving K > 0 reversed spins in the elec-
tronic ground state have a spin K+1/2. Under the particle-hole symmetry,
the spin of the quasiparticle-quasihole Hartree-Fock pair at = 1 is 2K+1,
where the positive integer K gives the number of reversed spins within the
Skyrmion (or anti-Skyrmion) and K = 0 corresponds to the spin-
1
2
quasi-
particle (or quasihole).
I next give two representative examples of Hartree-Fock calculations.
Firstly, microscopic Hartree-Fock calculations revealed that the absolute
19
Please note that the quantization of the total number of reversed spins (i.e., the fact
that the Skyrmion quantum number K is a positive integer) is not captured nor by this
variational wave function neither by the eld theoretic approach.
2.2. QHE FERROMAGNETISM AT = 1 73
value of the z-component of the spin is less than unity at the center of the
Skyrmion. The solid curve in Fig. (2.5) shows the Hartree-Fock magneti-
zation prole of the K = 3 Skyrmion
20
. This should be contrasted with
the result of the classical eld theory in which the spin is constrained to be
reversed at the center of the Skyrmion [dash-dotted curve in Fig. (2.5)].
Figure 2.5: The radial distribution of m
z
(r), from Hartree-Fock calculations
for a K = 3 Skyrmion, is shown by the solid curve. [Courtesy of M. Abol-
fath]. The dash-dotted curve is the result of the classical eld theory for the
ad hoc choice of /l
B
such as m
z
(r) = 0 at the same distance r, measured
from the center of the Skyrmion, as in the Hartree-Fock theory.
20
The number of reversed spins K is given by K =
1
4l
2
B
R
[1 m
z
(r)] d
2
r
1
2
.
74 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Secondly, the Hartree-Fock theory describes well the evolution of both Skyr-
mion spin and excitation energy with g. Using the Hartree-Fock approxima-
tion and taking into account the z-extent of the 2DES, Cooper [23] studied
Skyrmions at = 1 in the presence of a nite Zeeman coupling. Dening
a critical Zeeman energy by the value g at which the transition between
the charged spin-texture and the spin-
1
2
quasiparticle occurs, Cooper found
g
c
= 0.054 for bare Coulomb interactions [w
0
= 0, Fig. (2.6)].
Figure 2.6: Energy gap
t
1
[in units of e
2
/(l
B
)] for the creation of a Skyr-
mion anti-Skyrmion pair (solid curve) and total spin-1 quasiparticle
quasihole pair (dashed line) as a function of g = E
Z
/[e
2
/(l
B
)] from Hartree-
Fock calculations at = 1 and w
0
= 0 (bare Coulomb interactions). The
vertical arrow indicates g
c
= 0.054 above which a transition from Skyrmi-
ons to single spin-ip excitations is predicted. The spin of a single charge
excitation is shown in the inset. [Courtesy of N.R. Cooper]
2.2. QHE FERROMAGNETISM AT = 1 75
When the nite thickness of the conned state of electrons is introduced,
the critical Zeeman energy is reduced to g
c
0.046 (for sample M242,
w
0
/l
B
= 0.6) and g
c
0.047 (for sample M280, w
0
/l
B
= 0.5). Only the
result for M280 sample is shown here [Fig. (2.7)]. Notice that at each nite
g, when the nite thickness of the 2DES is included, the energy and the spin
of the Skyrmion are closer to those of the spin-
1
2
quasiparticle.
Figure 2.7: The solid curve is the energy gap
t
1
to create a Skyrmion
anti-Skyrmion pair [in units of E
C
= e
2
/(l
B
)] as a function of g = E
Z
/E
C
,
from Hartree-Fock calculations with w
0
/l
B
= 0.5 (sample M280) at = 1.
The dashed line is the Hartree-Fock result for the total spin-1 quasiparticle
quasihole pair. The vertical arrow indicates g
c
= 0.047 above which Skyr-
mions transform to single spin-ip excitations. The spin of a single charge
excitation is shown in the inset. [Courtesy of N.R. Cooper]
I should also mention here calculations of the Skyrmion excitation energy
which take into account both the nite thickness of the 2DES and the Lan-
76 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Table 2.1: The predicted g
c
below which the single spin ip excitation is
unstable to Skyrmion formation and the percentage decrease of the QHE
excitation gap at = 1 [(
t
1
)], relative to the value calculated at w
0
= 0
and r
s
= 0 [60, 76]. We recall that r
s
is the ratio of the Coulomb energy to
the cyclotron energy and w
0
parametrizes the z-extent of the 2DES.
g
c
(
t
1
) g
c
(
t
1
)
w
0
= 0 w
0
= l
B
r
s
= 0 0.061 0% 0.039 40%
r
s
= 1 0.052 30% 0.035 55%
dau level (LL) mixing. The LL mixing has the same qualitative eect as
the nite thickness of the 2DES on the stability of Skyrmions: it lowers
the Skyrmion excitation energy
21
and reduces g
c
. Interestingly, these cor-
rections do not work coherently [60, 76], i.e., the order in which these
corrections are included is important [Table (2.1)]. For example, when the
2DES is assumed innitely thin (w
0
= 0), the QHE excitation gap at = 1
is 30% smaller for r
s
1 compared to its r
s
= 0 value. A further 25%
reduction of the QHE excitation gap occurs, due to the inclusion of a nite
width of one magnetic length. On the other hand, considering that r
s
= 0,
with the introduction of a nite width of one magnetic length the energy
gap drops 40%. Then, the additional reduction of the energy gap due to
LL mixing is only 15% for r
s
1.
2.2.4 Electron spin polarization at = 1
Various thoughtful approaches have been used to study the T-dependence
of the electron spin polarization T at = 1 [19, 43, 50, 53, 87]. Our start-
ing point is T(T) calculated for a non-interacting 2DES with the chemical
potential in the middle of the Zeeman gap
T(T) = tanh
_
E
Z
/(4k
B
T)

. (2.43)
At high temperatures, according to Eq. (2.43), the 2DES behaves like a
Curie paramagnet, i.e., T(T) 1/T. A well-justied microscopic theory
21
In the absence of LL mixing, the particle-hole symmetry around = 1 ensures that
the energy required to create a Skyrmion is equal to that for creating an anti-Skyrmion.
When the LL mixing is included into calculations this equality can no longer be assumed.
2.2. QHE FERROMAGNETISM AT = 1 77
is the Hartree-Fock approximation, which represents the analog for QHE
systems of the band theory in itinerant electron magnetism, and yields
T(T) = tanh
_
E

HF
/(2k
B
T)

. (2.44)
Here E

HF
is the Hartree-Fock energy of spin-up electrons, measured from
the chemical potential (E

HF
= E

HF
= E
xc
/2). Because of the exchange-
enhanced spin splitting, T is much larger at nite T than in the non-
interacting case. Hartree-Fock approximation grossly overestimates the elec-
tron spin polarization: for particular parameters of Fig. (2.8), the theory
predicts that the 2DES is essentially 100% polarized below 10 K [53]. The
simplest many-body theory beyond the Hartree-Fock approximation which
reects the presence of the spin waves excitations in itinerant ferromagnets
is one which includes a self-energy insertion consisting of a ladder sum of
repeated interactions between electrons of one spin and holes of opposite
spin [53]. The spin-wave theory predicts a strong suppression of the elec-
tron spin polarization at much lower T because it includes the eects of
thermally excited spin waves. Calculations by Kasner and MacDonald [53],
for a static screening wavevector 0.01l
1
B
, are shown by the dash-dotted curve
in Fig. (2.8). The detailed T-dependence of T is as follows. At low T, the
reduction of T is dominated by the long-wavelength spin-wave contribution
which gives [1 +((T)T ln 1 e
E
Z
/(k
B
T)
], where ((T) depends weakly on
temperature. At higher temperatures, this suppression crosses over from an
an activated T-dependence [1((T)T exp E
Z
/(k
B
T)] for k
B
T < E
Z
, to
an approximately linear temperature dependence [1 ((T)T ln (k
B
T/E
Z
)]
for k
B
T > E
Z
. This linear T-dependence was interpreted as the analog of
the T
3/2
-Bloch law in 3D ferromagnets for QHE systems. Although the re-
sults shown in Fig. (2.8) are extremely insensitive to the choice of the screen-
ing wavevector, progress in the diagrammatic technique approach including
temperature and frequency-dependent screening has also been reported [50].
Kasner et al. [54] substantiated the spin wave theory, by appreciating the
role of the nite thickness of the 2DES [54]. The electron spin polarization
decreases with T more quickly when the nite width of the 2DES is taken
into account. A complementary method to compute T(T) was suggested by
Read and Sachdev [87], in the context of a continuum quantum eld theory
for the QHE ferromagnet at = 1. Their approach belongs to the class of the
so-called large-N methods
22
. By considering
s
as a parameter, they used
22
In their work, the method under consideration found double application as it permits
also to calculate the nuclear spin-lattice relaxation rate [see Ref. [99] and Sec. (2.3.5)
below].
78 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Figure 2.8: The T-dependence of T at = 1 and g = E
Z
/E
C
=
2.2 K/136 K = 0.016. The dashed curve illustrates T(T) for non-interacting
electrons [Eq. (2.43)]. The dotted curve represents the spin-
1
2
Brillouin func-
tion. The dash-dotted line shows the result of the spin-wave theory [53].
SU(N) and O(N) symmetries in the N = limit and exploit dimensional
analysis to compute T(T) at = 1. The
s
= 0 limit of the SU(N = )
model yield the s =
1
2
Brillouin function, which is displayed in Fig. (2.8) by
the dotted curve. The
s
= 0 limit of the O(N = ) model yield the s = 1
Brillouin function (not shown here). Working along the lines suggested by
Read and Sachdev [87], Green [43] obtained results for T(T) at = 1 in a
weakly disordered QHE ferromagnet and showed that the eect of disorder
is to reduce
s
, i.e., T decays more quickly with increasing T. Finally, we
wish to stress out a feature shared by all calculations: the expected T(T)
at = 1 for a disorder-free 2DES, lies always above the lower bound set by
the Eq. (2.43) [dashed curve in Fig. (2.8)]. That is, the predicted T(T) is
essentially saturated at very low-T (T 1 at T = 0.1 K).
2.3. QHE FERROMAGNETISM NEAR = 1 79
2.3 QHE ferromagnetism near = 1
Knowledge of the ground state precisely at = 1 and its low-energy exci-
tations is essentially simpler than the knowledge of the ground state near
= 1. The ground state near = 1 can be only approximately determined
because the QHE ferromagnet is non-colinear
23
. Colinear order means that
the electronic spins are all aligned, parallel or antiparallel, to the same di-
rection everywhere in space; non-colinear order means that the direction
of magnetization varies from point to point in space. Magnetic order near
= 1 is no longer colinear forasmuch the ferromagnetic ground state was
diluted with magnetic impurities (Skyrmions) that altered the balance of ex-
change interactions. The electron spin stiness is no longer the only relevant
energy scale near = 1.
2.3.1 Electron spin polarization near = 1
Without referring to any model, we can immediately settle a fundamental
prediction about the electronic ground state at ,= 1. This prediction,
which concerns the electron spin polarization, naturally emerges from the
properties of QHE Skyrmions we already presented. For lling factors 1
(or 1) there will be a nite density of Skyrmions (or anti-Skyrmions) in
the electronic ground state, all with the same charge e (or e) and carrying
the same number of reversed spins K. We assume here that there is electron-
hole symmetry around = 1. One can easily prove that
T() =
S
z
()
^/2
=
_
1 2K(1 )/, for 1,
1 2(K + 1)( 1)/, for 1.
In this picture, the lowest spin-up Landau level is completely lled at = 1
[^ = ^

= An

]. The 2DES is fully polarized at zero temperature T = 1


[S
z
( = 1) = ^/2]. Reduction of the magnetic eld so that ^ = ^

+ 1,
creates one Skyrmion, S
z
( 1) = ^

/2 (K +1/2), and the electron spin


polarization is reduced. Alternatively, increasing the magnetic eld so that
^ = ^

1, creates one anti-Skyrmion, S


z
( 1) = ^

/2(K+1/2), which
also reduces the electron spin polarization. The above equation suggests
that a simple way to determine K would be to measure the electron spin
23
Non-colinearity can generally be traced to competing interactions. For example, imag-
ine a triangular lattice with classical spins and with two types of exchange interactions.
Several structures (ferromagnetic, antiferromagnetic, and spiral) may occur, depending on
the sign and magnitude of the exchange coupling constants.
80 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
Figure 2.9: The -dependence of the theoretical spin polarization for non-
interacting electrons is shown by the dashed curve. The dotted curve repre-
sents the -dependence of T for [ 1[ 0.1, when non-interacting K = 3
Skyrmions are accommodated in the electronic ground state [36]. The solid
curve displays T() for a square Skyrmion lattice at g = 0.015 [14].
polarization as a function of in the vicinity of = 1. The predicted
reduction of T for K = 3 Skyrmions at ,= 1 is shown by the dotted line in
Fig. (2.9). The rapid decrease in the electron spin polarization is expected to
persist only in the QHE plateau [1[ 0.1. Brey et al. [14] obtained results
for T() using a Hartree-Fock approximation [solid curve in Fig. (2.9)].
These authors noted that the slope of the spin polarization curve decreases
in magnitude away from = 1, indicating that Skyrmions/anti-Skyrmions
shrink into single reversed-spin quasiparticles as they become more dense.
Recently, Nederveen and Nazarov [82] pointed out that the presence of a
nite density of Skyrmions anti-Skyrmions in the QHE plateau at = 1
would manifest as a rounding of the spin polarization peak. We also note
2.3. QHE FERROMAGNETISM NEAR = 1 81
that T() obtained for K = 0 corresponds to the independent (spin-ip)
electron model
T() =
_
1, for 1,
2

1, for 1 2.
When = 2, there is an equal number of spin-up and spin-down electrons
and T = 0. The free electron picture is displayed by the dashed curve in
Fig. (2.9). If the non-interacting electron picture is a good rst approxima-
tion for the 2DES in the quantum limit, it should describe well the measured
T() in the limit of T 0 and very large g. These considerations will serve
as a main frame for the discussion of our NMR data.
How the presence of Skyrmions near = 1 is reected in heat capacity
and NMR experiments, is the theme of the following two chapters. The cen-
tral aspect is that for lling factors away from = 1 some of the electron spin
uctuations are orders of magnitude lower in frequency than the electronic
Zeeman splitting. These nearly gapless modes allow a strong coupling to the
QWs nuclei, observed as a dramatic enhancement of the nuclear spin-lattice
relaxation rate near = 1 in both heat capacity and NMR experiments. In
a disorder-free 2DES, to understand this behavior, we must conceive how
Skyrmion dynamics involves low frequency spin uctuations.
To understand the properties of spin excitations of the QHE ferromagnet
near = 1 is merely the most arduous task. Our additional labours could be
distilled into the following questions: What are the symmetries of the ground
state containing Skyrmions? What is the nature of the multi-Skyrmion
system at T = 0? It is a solid or liquid? What is the character of phase
transitions, if any? These questions have an answer, quite complicated it
is true, and in part still very hypothetical, but yet deserving respect as it
provides an explanation of important experimental facts.
2.3.2 Quantum treatment of a Skyrmion
Returning to the analysis of one Skyrmion [parametrized in Eq. (2.18)],
we recall that there are two degeneracies in the classical eld theory. The
energy E
1
Sk
does not depend on (1) position of the Skyrmion and (2) in-
plane orientation of components of the spin. These degeneracies are sources
of nearly gapless excitations which license the nuclear spin-lattice coupling.
The way to see this is to treat the Skyrmion quantum mechanically [39, 97].
First, as the Skyrmion has a distinguishable location, a translational
degree of freedom [r
0
(t)] is added to the system. The spin eld of a moving
Skyrmion is dened as m(r, t) = m[r r
0
(t)] and the rate of change of the
82 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
local spin orientation is
m

(r, t) = r

[r r
0
(t)].
In the rst order approximation, the eective Lagrangian for the transla-
tional degree of freedom reduces to
L
m
= s
_
m

(r, t)/

[m(r, t)] n(r) d


2
r, (2.45)
where s =
1
2
and we have taken into account our new-found knowledge that
the charge density is non-uniform. As discussed in Sec. (2.1.3), this La-
grangian is equivalent to that of a massless charge e moving in a uniform
magnetic eld B. The addition to the Lagrangian of an inertial term pro-
portional to [ m[
2
would induce a mass for the Skyrmion, and permit it to
make cyclotron orbits. Therefore, like the massive degeneracy associated to
the 2D dynamics of a charge in the presence of a uniform magnetic eld, the
eigenstates of the Hamiltonian describing a moving Skyrmion are degenerate
in energy and therefore capable of relaxing the nuclei.
Secondly, since the Skyrmion has a distinguishable orientation, the spin
symmetry is broken with respect to rotations of the non-colinear spin con-
guration around the magnetic-eld direction and a rotational U(1) degree
of freedom [] is introduced in the system
24
. The eective Lagrangian for
the rotational degree of freedom is
L

= K

+

2
2U

2
, (2.46)
where K and U are phenomenological parameters [39]. The corresponding
Hamiltonian is
H

=
U
2
(i

K)
2
. (2.47)
Its eigenfunctions are
Sk
m
() =
1

2
e
im
and its eigenvalues are
Sk
m
=
U
2
(m K)
2
. The expectation value of the canonical angular momentum
operator conjugate to , i

, is K. Thus, for each Skyrmion, K and are


canonically conjugate. Skyrmions carry an integer-valued quantum number
K, which physical meaning is the number of overturned spins. The actual
orientation angle is completely uncertain because [
Sk
m
()[
2
= 1. This im-
plies that the U(1) rotational symmetry, broken in the classical solution, is
24
The global U(1) symmetry of the colinear ferromagnet means that the ground state
is invariant (up to a phase) under the transformation dened by R() = e

S
z
.
2.3. QHE FERROMAGNETISM NEAR = 1 83
restored in the quantum solution because of quantum uctuations in the
coordinate . However, in the thermodynamic limit of an innite number of
Skyrmions coupled together, it is possible for the U(1) rotational broken sym-
metry to survive quantum uctuations. If this happens, then the excitation
spectrum is gapless and the nuclear spin-lattice relaxation is allowed.
2.3.3 Skyrmion lattices at zero temperature
Moving away from = 1, more and more Skyrmions are present in the
electronic ground state and their interaction becomes important. Theoreti-
cal calculations suggest that the ground state of the 2DES near = 1 is a
Skyrmion crystal [14, 24]. The qualitative understanding of various types
of Skyrmion lattices and multi-Skyrmion ground states has drifted into the
view that there are two competing interactions in the system: (1) a repul-
sive, long-range Coulomb part (because Skyrmions are charged) and (2) a
short-range contribution related to the U(1) degrees of freedom. The latter
term, the so-called magnetic interaction, favors antiferromagnetic arrange-
ment of U(1) spins. Since the magnetic interaction is of short range, the
Coulomb interaction gives the dominant energy scale for small Skyrmion
densities and a triangular lattice is preferred at very small values of [ 1[.
The orientational degree of freedom of the Skyrmion is then frustrated
25
.
In the triangular phase the orientation angle diers by 120

between neigh-
boring Skyrmions. At higher Skyrmion densities, the energy gained from
the XY antiferromagnetic ordering outweighs the Coulomb energy cost and
a square lattice structure is rather stabilized. The orientation angle is then
shifted by 180

between neighboring Skyrmions. Further lattice types may


also be possible at intermediate Skyrmion densities
26
. The main theoretical
prediction is that the non-colinear ground state breaks both translational
and U(1) rotational symmetry [24]. Associated to these broken symmetries
(by virtue of Goldstones theorem) there are two Goldstone modes, namely,
lattice vibrations and XY spin waves. The Goldstone mode associated with
the broken translational symmetry is just the ordinary magneto-phonon
mode of the crystal and has a power-law dispersion. More precisely, the
dispersion relation is E(q) =
_
2 l
B
e
2
/
_
1/2
_
ql
B
_
3/2
, where is the shear
modulus and q is the magnitude of the wavevector q. At long wavelengths,
25
The key distinction between ferromagnetic and antiferromagnetic interactions is that
the latter are subject to topological frustration (one can devise networks where all the
exchange bonds can not be satised simultaneously), whereas the former are not.
26
These considerations pertain to a (mean-eld) classical model. At the highest Skyr-
mion densities, quantum uctuations are expected to give rise to a quantum-melted state.
84 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
the magneto-phonon mode is essentially decoupled from the XY spin-wave
mode. Thus, ignoring the positional degrees of freedom, the ground state
theory of the QHE ferromagnet near = 1 reduces to the subtle physics of
U(1) degrees of freedom, on which we now focus. A phenomenological XY
lattice antiferromagnetic model was introduced in Ref. [24] for Skyrmion
crystals
H
multiSkyrmion
=

U
2

j
(i

j
K)
2

ij)
cos (
i

j
), (2.48)
where single indices label sites on the lattice and

ij)
means that the sum
is taken over all pairs of nearest neighbors ij). Here

U accounts for the
energy cost when the i

j
s eigenvalue deviates from the optimal value
K. The stiness to relative rotations of neighboring Skyrmions is

J < 0,
consistent with the antiferromagnetic XY order. This model is equivalent to
the model of bosons hopping on a lattice
27
, in which the boson number on
the ith site is mapped to the number of ipped spins in the ith Skyrmion. In
the boson language, the orientation-dependent interaction term corresponds
to the boson hopping and favors the long range phase coherence of bosons,
and

U is the strength of the on-site interaction between bosons. For

U

J,
the system is an insulator and there is an excitation gap of the order of

U.
For

U

J, the system is a Bose superuid, the U(1) rotational symmetry
is not restored by quantum uctuations, and the corresponding Goldstone
mode has a linear dispersion at small wavevectors
E(q) = (

U

J)
1/2
q a, (2.49)
where a is the Skyrmion lattice constant. If this situation occurs, then an
excitation gap is not produced and the nuclear spin-lattice relaxation rate is
dramatically enhanced
28
. The parameters

J and

U can be xed by tting
the microscopic Hartree-Fock calculations. Using the boson language, the
boson interaction parameter

U is given by the inverse boson compressibility

U =
1
2
_
K( g, ))
g
_
1
. (2.50)
The boson hopping parameter can be extracted then from the collective
mode spectrum. These considerations allow an estimate of the nite range
27
M.P.A. Fisher, P.B. Weichman, G. Grinstein, and D.S. Fisher, Boson Localization and
the Superuid-Insulator Transition, Phys. Rev. B 40, 546 (1989).
28
In the Bose superuid phase, the number of overturned spin is uncertain and so is
the Zeeman energy cost to ip an electron spin.
2.3. QHE FERROMAGNETISM NEAR = 1 85
in the g- phase diagram where it is possible for the 2DES to support gap-
less elementary excitations [Fig. (2.10)]. It is worth noting here that the
slight distortion in static positions of Skyrmions, in response to a disorder
potential, will give a small random contribution to the stiness

J. This
randomness is expected to produce a region of Bose-glass phase.
Figure 2.10: Qualitative, zero-temperature g- phase diagram for Skyrmion
crystal states [24]. The lled circles indicate the values of K along the square
Skyrmion lattice region boundary. Non-colinear magnetic order survives
quantum uctuations in the region approximately indicated by the shading.
2.3.4 Skyrmion lattices at nite temperatures
It is generally accepted that the qualitative distinction between dierent
multi-Skyrmion magnetic phases at T = 0 will survive at nite T. If Skyr-
mion positions were xed to the ideal lattice at all temperatures, the above
86 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
discussed XY lattice antiferromagnetic model would capture the long wave-
length physics. In the Skyrmion lattice, however, the positions are not xed
and the lattice itself can melt. In the melted phase, long length-scale corre-
lations vanish, but we expect that short length-scale correlations to contain
crystal-like features. Therefore, even at temperatures over the melting tem-
perature, the multi-Skyrmion system could be analyzed in quite the same
spirit as at T = 0. For example, above the lattice melting temperature,
Goldstone modes of the 2DES will be overdamped modes derived from zero-
temperature coherent modes.
At high T, the multi-Skyrmion system loses its quasi-long range order
in position and orientation. From a phase transition viewpoint, one expects
two Kosterlitz-Thouless phase transitions as T is lowered. A rough upper
bound estimate for the critical temperature of the Kosterlitz-Thouless phase
transition associated with the loss of magnetic quasi-long range order is [24]
T
magnetic
KT


2k
B

s
1 K, (2.51)
where

U = 0 and

J
s
. The theoretical estimate for the critical tempera-
ture of the solid-to-liquid Kosterlitz-Thouless phase transition is [24]
T
melting
KT
=
a
2

4
0.1 K, (2.52)
Thus, the magnetic disordering transition is expected to occur at a higher T
than the lattice melting transition
29
. However, despite the fact the collec-
tive modes are largely decoupled, the topological excitations (vortices and
dislocations) may be coupled, leading to an interplay between magnetic and
melting transitions. Dislocation-driven simultaneous transitions and vortex-
driven simultaneous transitions are two physically possible scenarios [100].
Therefore we can surmise that near = 1, at least one Kosterlitz-Thouless
phase transition is expected as T is lowered.
2.3.5 Skyrmions and the nuclear spin-lattice relaxation rate
If one is to pick out a single experimental feature that epitomizes Skyrmion
vita this would be the nuclear spin-lattice relaxation rate T
1
1
. Sharp struc-
tures in T
1
1
as a function of T would be strong indications for phase transi-
tions in the multi-Skyrmion system
30
. I discuss here two signicant theoreti-
29
The Skyrmion lattice spacing is about 10
3
larger than the GaAs lattice spacing.
30
The nuclear spin-lattice relaxation in GaAs QWs and the general modelling of T
1
1
in QHE systems will be reviewed in Chapter 4.
2.3. QHE FERROMAGNETISM NEAR = 1 87
cal achievements concerning T
1
1
(T, ). I shall start with the bosons hopping
model [24] and then proceed to a quantum non-linear eld model [44, 87].
Calculations by Cote et al. [24] in the ordered phase of the multi-
Skyrmion system predict a remarkable evolution of T
1
1
with T and . Ac-
cording to this theory, when the Skyrmion lattice is pinned by the disorder,
the magneto-phonon spectrum is gapped and does not contribute to T
1
1
.
But the XY spin-wave spectrum do contribute to the nuclear spin-lattice
relaxation rate, and in the limit of zero temperature, T
1
1
has a linear rela-
tionship with both [ 1[ and T. It is to be noted that, taking into account
the combined eect of disorder and interactions, the electron spin dynamics
was theoretically studied in the past (well before ideas on QHE Skyrmi-
ons were put forward) and quantitative predictions for T
1
1
were obtained
around = 1 [5]. These calculations emphasized the importance of disorder
and electron-electron interactions to the nuclear spin-lattice relaxation rate.
Of paramount importance, they do not predict a slowing down of the nuclear
spin-lattice relaxation near = 1 [dotted and solid curves in Fig. (2.11)].
This is in sharp contrast with the bosons hopping model prediction that
T
1
1
0 when [ 1[ 0 [dashed lines in Fig. (2.11)]. We should also
remark that recent theoretical studies questioned how good the predicted
linear T-dependence actually is. We now turn to this aspect.
A weighty attack on the problem was opened by Read and Sachdev [87];
considering the scaling behavior of 2D Heisenberg models, these authors
succeeded in determining the T-dependence of the T
1
1
at = 1. Pursuing
the argument further, Green [44] pointed out the quantum critical nature of
uctuations of the Skyrmion lattice and arrived at similar results near = 1.
The dierent scaling regimes of 2D Heisenberg models could be understood
as follows. Recall that, according to the Mermin-Wagner theorem, when a
spin-wave expansion is attempted for a ferromagnet at or below its criti-
cal dimension, the occupation of low-frequency modes is found to diverge.
However, the constraint, xing the magnitude of the local spin, restricts this
divergence. The interplay between the divergence and constraint give rise
to a nite-T correlation length. The dynamics of the ferromagnet is very
dierent on the length scales greater or less than the correlation length. On
length scales less than the correlation length the ground state is ordered
and uctuations are purely classical, i.e., the characteristic frequency of
spin uctuations is much smaller than T. Accordingly, this regime is called
renormalized classical and occurs at T
s
. Here renormalized means
that the T = 0 parameters such as the spin stiness are renormalized com-
pared to their mean-eld values as a result of quantum uctuations at short
wavelengths. At length scales greater than the correlation length the system
88 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
is disordered and uctuations are overdamped. This is the quantum critical
regime and it is the high-T scaling regime (T
s
) in the phase diagram.
The crossover between renormalized classical and quantum critical regimes
occurs at T
s
.
Figure 2.11: The dashed line pictorially describes the -dependence of T
1
1
in the bosons hopping model [24]. Calculations by Antoniou and Mac-
Donald [5] are illustrated for two dierent values of the ratio between the
exchange-enhanced Zeeman splitting (E
Z
+E
xc
) and the Landau level width
. The ratio (E
Z
+E
xc
)/ is either 0.85 (solid curve) or 0.15 (dotted curve).
In short, we expect the following T-dependence of T
1
1
in the QHE plateau
at = 1. Starting from the high temperatures, T
1
1
is constant with T in
the quantum critical region. If the temperature is decreased, the system
crosses to the renormalized classical regime. In this regime T
1
1
increases
exponentially fast with decreasing T. At even lower temperatures, precisely
at = 1, T
1
1
is expected to display an activated behavior. This should be
contrasted with the T
1
1
T law expected near = 1 at T 0. Finally, at
2.4. INTERPRETATION OF THE = 1 QHE EXCITATION GAP 89
T = 0 the nuclear spin-lattice relaxation rate is zero.
2.4 Interpretation of the = 1 QHE excitation gap
Looking back at this edice, we may briey examine its ground-plan. The
2DES at = 1 has no genuine charge excitations. The number of reversed
spins within a charge excitation at = 1 varies with g. Provided that g
is small, the low-lying charged elementary excitations are Skyrmions, en-
compassing many spin reversals. At large g, they are single spin ips. Ac-
cordingly, the meaning of the QHE excitation gap at = 1, probed by the
magnetotransport, varies with g. At small g, it represents the energy cost
for the creation of a Skyrmion anti-Skyrmion pair, whereas at large g it
represents the energy cost for the creation of a widely separated quasiparticle
quasihole pair.
Investigations of the Skyrmion physics by magnetotransport were rst
reported by Schmeller et al. [89]. These authors used tilted magnetic elds
to tune g and measured the QHE excitation gap at = 1 (
1
) on single-
QW samples. Noteworthy, in the tilted-magnetic eld technique g increases
as the tilt angle increases and its minimal value is recorded at = 0

.
Henceforth, the tilted-magnetic eld technique is particularly suitable for
Skyrmion studies at large g. A series of measurements at = 1 concentrated
on the range of small g. To achieve the g 0 regime several ingenious
experimental schemes have been proposed [75, 92]. The principal task of
magneto-transport experiments [75, 89, 92] is the direct measurement of the
number of reversed spins within a Skyrmion. The total spin of the thermally
activated Skyrmion anti-Skyrmion pair at = 1 can be extracted from the
change in
1
produced by tilting the total magnetic eld B. It is assumed
that an in-plane magnetic eld component couples only to the spin degrees
of freedom, while the perpendicular magnetic eld component B

controls
the orbital dynamics. This implies that the Zeeman contribution to the
energy gap
1
for creating a pair of elementary excitations with total spin
2K + 1 (in units of ) from the fully polarized ground state is
31
(2K + 1)[g

[
B
B =
1

0
1
. (2.53)
The term
0
1
is the contribution to the gap from all non-Zeeman sources and,
in this model, depends only upon B

. It follows that
1
/B (evaluated
31
The traditional (i.e., pre-Skyrmion) view of the = 1 QHE excitation gap have
assumed K = 0.
90 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
at = 1, where B

is xed) is just (2K + 1)[g

[
B
. Since the g

and
B
are known, measuring
1
/B determines the quantum number K.
Figure 2.12: Measured QHE excitation gap at = 1 (
1
) vs Zeeman energy,
both in units of e
2
/(l
B
) (sample M280). The corresponding tilt angles are
indicated on top axis. The dotted and dashed lines correspond to 2K+1 = 9
and 2K +1 = 1, respectively. The uncertainty in the measured
1
is 6%.
I present here QHE excitation gap measurements at = 1 in tilted mag-
netic elds, carried out on M280 sample. To our knowledge, this study
was performed up to the largest yet reported g 0.05. The value of the
QHE excitation gap at = 1 was determined from the T-dependence of
the longitudinal resistance in the thermally activated regime, as described
in the precedent chapter. The evolution of
1
with tilt angle is presented
2.4. THE = 1 QHE EXCITATION GAP 91
in Fig. (2.12) by plotting
1
vs Zeeman energy, both expressed in units of
e
2
/(l
B
). Figure (2.7) shows the calculated Hartree-Fock energy gap for the
creation of a widely separated Skyrmion anti-Skyrmion pair as a function
of g for sample M280. This calculation take into account the nite thickness
of the electron layers. The relevant parameter is w
0
/l
B
= 0.5, where w
0
represents the rms width of the self-consistently calculated subband wave
function in each QW. Similar to results for single QWs [89], there is an
overall qualitative agreement between the measured and calculated gaps in
Fig. (2.7). At = 0

,
1
= 20 K, comparable to the measured
1
in high-
quality conventional single-layer 2DESs. Note that
1
( = 0

) is only 16%
of the theoretical value for an ideally thin, disorder-free 2DES, and with no
Landau level mixing. The discrepancy originates from the inadequacies of
this simplistic model for the 2DES. It is obvious from Fig. (2.7) that the
nite z-extent of 2D electron layers causes a dramatic reduction of the cal-
culated
1
, reducing the discrepancy between theory and experiment to a
factor of 4. Even with the inclusion of both disorder and LL mixing, the
calculated excitation gap remains 2 times larger than the experimental one.
More important, assuming that the slope 2K + 1 =
1
/([g

[
B
B) gives
the spin of a pair of charged excitations at = 1, both experiment and
theory give K

= 4 for g 0.012 and K

= 0 in the limit of large g. The con-
tent of Figs. (2.7) and (2.12) demonstrates very satisfactorily a convergence
between the theoretical description of Skyrmions and their experimentally
determined properties. However, given the experimental uncertainty in the
measured gaps, the absence of quantitative agreement with theoretical val-
ues and the fact that
1
is expected to slowly approach the single spin-ip
dependence (2K + 1 = 1), prohibit an accurate determination of g
c
based
on transport measurements.
One of the objectives of heat capacity and NMR experiments to be de-
scribed in the following two chapters is to clarify the problem of the range
of g over which Skyrmions are the relevant excitations of the = 1 QHE
ground state, by exploring the large- g limit. In brief, tilted-magnetic eld
heat capacity measurements on M280 sample show evidence for a transition
from Skyrmions to single spin-ip excitations at g
c
0.04. Our NMR data
in tilted-magnetic elds, collected on M242 sample, reveal the presence of
Skyrmions in the electronic ground state for g 0.022 and their absence for
g 0.037. These experimental values are in good agreement with Hartree-
Fock predictions which take into account the nite thickness of the 2DES:
g
c
0.046 (for M242 sample) and g
c
0.047 (for M280 sample).
92 CHAPTER 2. QUANTUM HALL EFFECT SKYRMIONS
2.5 Summary
In this chapter we have studied the QHE ferromagnet at = 1 and intro-
duced the concept of QHE Skyrmions [95]. The Skyrmion is a spin texture
with spins down at the origin but up at innity; at intermediate distances
the spins lie in the XY plane and have a vortex like conguration. The
importance of Skyrmions for QHE systems arises from the discovery that
they describe the lowest energy charged-excitations of the QHE ferromag-
nets. Therefore, Skyrmions dominate many of the low-T properties of these
systems, and we have paid a particular attention to the T-dependence of the
electron spin polarization at = 1.
Properties of this excitation, such as its energy and spatial extent, are
determined by the ratio g of the single-particle Zeeman energy, which limits
the number of spin ips within an excitation, to the Coulomb energy which
favors local ferromagnetic ordering. The spin of the thermally activated
Skyrmion anti-Skyrmion pairs was deduced from the g dependence of the
QHE excitation gap at = 1. Our results reveal that the spin of a pair of
charged excitations at = 1 is 9 ( g = 0.012 in sample M280). Although
consistent with the measured Skyrmion size, Hartree-Fock calculations [23],
even after Landau level mixing and nite-thickness corrections, overestimate
the QHE excitation gap at = 1 over the entire range of our data.
We also addressed the properties of QHE ferromagnets near = 1,
including the -dependence of the low-T electron spin polarization. The
ground state of the 2DES at lling factors slightly away from = 1 is a
crystal of Skyrmions. This means that the system has non-colinear order.
Associated with this is a new U(1) degree of freedom in which the spins
rotate about the magnetic eld direction. Because Skyrmions contain spins
lying in the XY plane, such rotations do represent new states. Since the
rotation is about the magnetic eld direction, this collective mode does not
have a Zeeman gap, but rather is a gapless Goldstone mode. Therefore,
the 2DES strongly couples to the nuclear spin system and the nuclear spin-
lattice relaxation rate at 1 and 1 increases by a factor of 10
3
over
the zero magnetic eld value. We shall use the formalism developed in this
chapter to analyze the following heat capacity and NMR experiments.
Chapter 3
Heat Capacity Evidence for
Skyrmions
Pentru ca nsusi Domnul, ntru porunca, la glasul Arhanghelului
si ntru trambit a lui Dumnezeu, Se va pogor din cer, si cei mort i
ntru Hristos vor nvia ntai. Dupa aceea, noi cei vii, care vom
ramas, vom rapit i, mpreuna cu ei, n nori, ca sa ntampinam
pe Domnul n vazduh, si asa pururea vom cu Domnul.
Sf. Apostol Pavel
3.1 Introduction
3.1.1 Small-sample calorimetry
Hardly one can nd elsewhere in solid-state physics the need for high-
resolution small-sample calorimetry more imperatively expressed than in
the area of two-dimensional electron systems (2DESs). I have in mind not
only the low strength of the measured signal, resulting from the tiny mass of
the specimen and the small number of electrons, but also the heinous regime
of low temperatures (T 100 mK) and high magnetic elds (B 10 T), fre-
quently of principal interest. The low-temperature small-sample calorime-
try in high magnetic elds has seen a number of breakthroughs both in
design concept and instrumentation over the years before the discovery of
the quantum Hall eect (QHE). In the experiments which form the object
93
94 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Table 3.1: Landmark papers describing calorimetric techniques for small-
sample heat capacity measurements. Denitions are given in the text.
Year Denition Method
1963 C = Pt/T adiabatic heat-pulse
a
1968 C = P/(

2
H
T
ac
) steady-state, ac-temperature
b
1972 C = P/T weak-link time-constant
c
1977 C = (P/T)[1 exp (t/)] weak-link heat-pulse
d
a
Morin and Maita, Ref. [81].
b
Sullivan and Seidel, Ref. [98].
c
Bachmann et al., Ref. [6].
d
Fagaly and Bohn, Ref. [34].
of this chapter we have employed three calorimetric techniques. Inuential
papers describing these methods are listed in Table (3.1). The adiabatic
heat-pulse technique make use of the classical denition of the heat capac-
ity: C = Pt/T. The temperature of the sample rises by an amount T
when the sample (thermally isolated from its surroundings) is subjected to
a small heat pulse of power P and duration t. In the steady-state ac-
temperature calorimetry, an ac heating signal [P cos (
H
t)] is applied at the
frequency f
H
=
H
/(2) to the sample, which is thermally connected to its
surroundings. In the rst-order approximation, the rms amplitude of the re-
sulting temperature oscillations T
ac
is simply related to the heat capacity:
C = P/(

2
H
T
ac
). Another appropriate technique for high-resolution
small-sample heat capacity measurements, referred to as weak-link time-
constant method, is also contingent upon the thermal link connecting the
sample to the heat sink. A constant power P is applied to the sample so that
the steady-state temperature of the sample rises by an amount T. Then,
the power is turned o and the temperature of the sample decays back to
the sink temperature with a characteristic time constant . The total heat
capacity
1
is given by C = P/T.
While these preliminary remarks may suce to determine the principles
axed to calorimetric methods used in this study, they do not tell the whole
story. They do not reect the fact that, frequently, in real calorimetric
systems two time constants characterize the thermal dynamics of the sample:

ext
describes the thermal coupling between the sample and its surroundings,
1
If the pulse duration is short, a suitable formula is C = (P/T)[1 exp (t/)].
3.1. INTRODUCTION 95
whereas
int
describes the thermal relaxation within the sample. Simple
reection upon the most common experimental situations is sucient to
clarify the physical meaning of
ext
and
int
. For a rst orientation as well
as for the later considerations, lets consider relaxation-time heat capacity
measurements. In this case the sample is connected to a heat sink through
a heat leak. We assume that the heat leak has a nite thermal conductance
K and zero heat capacity, the heat capacity of the sample is C, and the heat
sink has a heat capacity C
s
(C C
s
). To start an experiment, the sample
is heated by a constant power P so that the temperature of the sample
rises by an amount T. Then, after the steady-state has been attained,
the power is turned o. If the sample is a perfect thermal conductor, the
temperature of the sample will decay back to the sink temperature with a
single characteristic time constant (alias
ext
). One can easily see that C
is obtained from two independent measurements: (1) and (2) K = P/T.
That is, the external time constant is determined by C and K: = C/K.
But when the sample is, for example, a poor thermal conductor, one
has to carefully consider the heat ow through it, and to take into ac-
count its internal relaxation time
int
. The internal thermal relaxation might
lead to a markedly non-exponential temperature decay and the conventional
relaxation-time method (described above) need to be corrected
2
. Lets be
more specic and suppose that, in the example under consideration, the heat
is carried by the phonons. Then,
int
can be estimated from the thermal
diusivity data:
int
L
2

/D

. Here D

=
1
3
v

is the thermal diusivity,


v

is the speed of sound,

is the phonon mean free path, and L

is the
largest dimension of the sample [115]. While it is fairly evident that for
successful heat capacity experiments
int
should be smaller than
ext
, it is
less obvious that the experimental values of
ext
and
int
will force upon us
the adoption of one calorimetric technique or another [Table (3.2)].
The above example does not encode all forms of
int
. Low temperatures
combined with high magnetic elds challenge us with a truly dangerous
trap: the nuclei might couple to the degrees of freedom whose heat capacity
we would like to measure (phonons or/and electrons). Traditionally, this
coupling is described by the nuclear spin-lattice relaxation time T
1
. When
heat is put into the sample, T
1
plays the role of an internal time constant,
i.e., it describes the establishment of the thermal equilibrium between the
nuclear spin system (NSS) and the rest of the sample (the lattice)
3
. Since
2
A notorious experimental signature of the
int
-eect is an overshoot in the samples
temperature when heat is turned on suddenly.
3
In the nuclear magnetic resonance terminology (e.g., nuclear spin-lattice relaxation
time), the lattice denotes all degrees of freedom of the sample excluding the nuclei.
96 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Table 3.2: Calorimetric methods for heat capacity measurements in the QHE
regime. In principle, provided that more restrictive experimental conditions
are fullled [40], any of these methods can be used for accurate ( 1%) heat
capacity measurements. In our experiments the uncertainty is 10%.
Method
ext
Condition
adiabatic heat-pulse 10
4
10
5
s
ext
> t >
int
weak-link time-constant 10 10
3
s
ext
>
int
steady-state, ac-temperature 0.01 1 s
ext
> 1/
H
>
int
at low temperatures, T
1
could be very long,
int
-eects might appear in
relaxation-time heat capacity measurements. A main theme of this chapter
is the behavior of the heat capacity of multiple-quantum well (QW) samples
when the nuclei couple to the lattice in the experimental time-scale. The
model for the calorimetric system, the equations governing the heat transfer,
and the expressions for C and T
1
will be discussed in detail.
3.1.2 Thermal equilibrium in multiple-QW samples
Multiple-QW heterostructures are unique examples of multiple-subsystems
samples, as they are composed of the lattice (phonons)
4
, the two-
dimensional (2D) electron layers, and the NSS of QWs and barriers
[Fig. (3.1)]. Since we are able to perceive now the dangers that beset low-T
heat capacity experiments on multiple-subsystems samples, before getting
down to the details, it is instructive to further remark on the process of
thermal equilibrium in our measured specimens [69].
Surprising thermal eects may happen in multiple-QW samples because
nuclei can be divided in two groups. The rst group is represented by the
nuclei located in the QWs. The second group is formed by the nuclei located
in the barriers (and the GaAs substrate). In most of the cases, the coupling
between the NSS and the lattice is so weak that, a fast refrigeration will leave
the nuclei at the starting T. A weird situation occurs when the coupling
between the QWs nuclei and the lattice is strong enough, but the coupling
between the barriers nuclei and the lattice is weak. In this case, a fast
4
The crystal lattice is the regular array of sites in the three-dimensional space where
the individual atoms are supposed to lie when the whole specimen is in its ground state.
3.1. INTRODUCTION 97
refrigeration will pull the QWs nuclei to low T while leaving the barriers
nuclei at the starting T. Finally, it is possible to refrigerate all nuclei in
thermal equilibrium with the lattice down to very low temperatures
5
.
Figure 3.1: Thermal diagram of a multiple-QW sample connected to a heat
sink. Subsystems are linked to each other by means of thermal resistances.
In general, heat capacity is a bulk property with contribution from all
components of a given thermodynamic system. Henceforth, depending on the
strength of the nuclear spin-lattice coupling at low T, one might measure:
(1) the lattice (phonons) and the electronic heat capacity, (2) the lattice
(phonons), the electronic, and the nuclear spin heat capacity of GaAs QWs,
(3) the lattice (phonons), the electronic, and the nuclear spin heat capacity
of GaAs QWs and barriers. An estimate of various heat capacities (to be
5
At low-T, it is advantageous in thermal studies to attribute a separate nuclear spin
temperature and heat capacity to each nuclear spin subsystem [Fig. (3.1)].
98 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
made shortly) will reveal that the heat capacity of a multiple-QW sample
might vary ve orders of magnitude! The magnitude of the heat capacity,
in turn, will critically constraint the choice of the experimental method
[Table (3.2)].
3.1.3 Samples for heat capacity experiments
For the purpose of heat capacity measurements, samples were prepared by
etching and polishing the material. The thinning of the excess bulk GaAs
substrate was necessary to remove the indium present on the back of the
samples from their molecular-beam-epitaxy block mount, and to improve
the 2DES signal-to-background ratio
6
. For heat capacity experiments on
M242 heterostructure we used a 7 7 mm
2
piece of the wafer thinned down
to 65 m. Heat capacity experiments on M280 heterostructure were
carried on a 7 10 mm
2
piece of the wafer which was thinned down to
160 m. Unfortunately, the expedient of using large area samples makes the
density not entirely homogenous within a 2D electron layer. Typical density
inhomogeneities across our measured specimens are 3%. The transverse
variation of density (from layer to layer) is 2%.
As stated before, in a multiple-QW heterostructure, various types of
heat capacities are encountered, each having its own T-dependence. We
estimate in the following the lattice (phonons) C
l
, the electronic C
e
, and
the nuclear C
n
heat capacities of our measured specimens. According to
the Debye model [115], C
l
shows a T
3
-dependence at low temperatures.
The phonon contribution to the specic heat of bulk GaAs amounts to c
l

(4
4
/5)3k
B
(8/A
3
c
)(1/
c
)(T/
D
)
3
= 26.9 (T/
D
)
3
[J/(gK)] for T <
D
/20, in
view of the crystal (
c
) and atomic density (8/A
3
c
) listed in Table (1.2). Here

D
345 K is the zero-T elastic Debye temperature of GaAs [12]. As-
suming that the Al
0.1
Ga
0.9
As and Al
0.3
Ga
0.7
As layers have the same phonon
specic heat as the bulk GaAs, the estimated low-temperature C
l
of our
samples (including the GaAs substrate) is
C
l
=
_
1.1 10
8
T
3
(J/K) for M242
3.9 10
8
T
3
(J/K) for M280,
where T is expressed in K. For non-interacting 2D electrons, in the limit
of low T such that k
B
T E
F
, C
e
has a linear T-dependence and it is
determined by the electronic density of states (DOS) at the Fermi level
T(E
F
): C
e
= (
2
/3)k
2
B
TT(E
F
) (J/K) per electron. Given the fact that
6
The thinning process of multiple-QW heterostructures is described in Ref. [110].
3.1. INTRODUCTION 99
m

/(
2
) = 0.24 10
14
[1/(m
2
K)] (including the spin degeneracy), the
estimated C
e
of our samples at zero magnetic eld is
C
e
=
_
6 10
12
T (J/K) for M242
8 10
12
T (J/K) for M280,
where T is expressed in K. When the sample is placed in a magnetic eld,
besides lattice and electronic heat capacities, there might be an excess heat
capacity due to nuclear spins. In the high-T limit, the Schottky nuclear heat
capacity of Ga and As atoms in the QWs (see Appendix A) is estimated at
C
QW
n
=
_
1.9 10
11
B
2
T
2
(J/K) for M242
3.3 10
11
B
2
T
2
(J/K) for M280,
whereas for the Schottky nuclear heat capacity of Ga and As atoms in the
barriers we obtain
C
B
n
=
_
14.1 10
11
B
2
T
2
(J/K) for M242
27.5 10
11
B
2
T
2
(J/K) for M280.
Here T is expressed in K and B is given in T. The Schottky nuclear heat
capacity of Ga and As atoms in the barriers is counted by making use of the
ratio between the width of the barriers and the QWs width. Estimates of
C
l
, C
e
, C
QW
n
, and C
B
n
(evaluated at T = 100 mK) are listed in Table (3.3).
At a glance, one can see that C might vary by up to ve orders of magnitude.
Table 3.3: Estimates of lattice (C
l
), electronic (C
e
), and nuclear (C
B
n
, C
QW
n
)
heat capacities (in pJ/K) for samples M242 and M280 at T = 0.1 K. Here
C
l
is the phonon contribution to the heat capacity of the entire sample
(including the GaAs substrate). The nuclear heat capacity of Ga and As
atoms is evaluated at B = 10 T. The thermal conductance of the heat leak
K
ext
(in pW/K) is estimated at T = 0.1 K (B = 0).
Sample C
l
C
e
C
QW
n
C
B
n
K
ext
M242
a
10 0.6 2 10
5
1 10
6
60
M280
b
40 0.8 3 10
5
3 10
6
60
a
The heat leak assumed here is a 60 m-diameter, 10 mm-long NbTi wire.
b
The heat leak considered here consists of forty 10 m-diameter GY-70
graphite bers of length 10 mm.
100 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.2: (Top) Photograph of the Bayotron rotating platform. (Bottom)
Experimental conguration for heat capacity measurements (sample M280).
3.1. INTRODUCTION 101
3.1.4 Heat capacity: Experimental details
Over a four-year period from 1992 to 1996, Prof. Vincent Bayot and Dr.
Eusebiu Grivei were heavily involved in bringing to fruition an apparatus
for heat capacity measurements at dilution refrigerator temperatures. At
the very beginning of my thesis they oered me a wonderful opportunity
to learn about low-T specic heat experiments. As the calorimetric system
sometimes requires manual adjustments of data collecting parameters (T,
B, time delays, ...), it stands to reason that a connection deepened between
myself and the low-T calorimetry.
Figure (3.2)(top) shows our rotating stage which permits in situ tilting
of the sample. The rotating platform is made from Stycast epoxy and it
accommodates a standard dual in-line multiple pin (DIP)-socket. If neces-
sary, the tilting stage is attached to the mixing chamber and it is thermally
anchored to the cold nger of the Bayotron. A carefully done wiring of
the tilting stage is an absolute must to ensure a well cooling of the sam-
ple and to minimize thermal drifts. This was accomplished with insulated
80 m-diameter copper wires which were thermally anchored by wrapping
them around the pins of the DIP-socket. The wires were further cemented
with GE 7031. The sample holder with an actual sample mounted for heat
capacity measurements is shown in Fig. (3.2)(bottom). The DIP-header
supporting the sample ts either into the DIP-socket xed onto the mixing
chamber or on the standard DIP-socket mounted on the tilting platform.
The sample could be positioned accurately and reproducibly in situ over a
range 0 90

within 0.1

, which represents the mechanical stability of the


rotating platform. The tilt angle is precisely determined by measuring the
Hall resistance on a 2DES device xed on the rotating stage.
For heat capacity measurements, the M280 sample was tilted in situ so
that an angle 0

77

formed between B and the normal to the sample


plane. For M242 sample, the experimental arrangement did not include the
rotating platform: C was measured while B was applied either perpendicular
to the 2DES plane ( = 0

) or at a xed angle of = 30 2

.
Thermometry
For the measurement of heat capacity, a heater and a thermometer should
be placed in good thermal contact with the sample. In our experiments,
the thermometer and the heater are carbon paint resistors deposited on the
substrate side of the sample. The carbon resistors were calibrated between
15 mK and 1 K against a RuO
2
resistance thermometer at B = 0.
102 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
It is important to note that carbon resistors were very stable, as they did
not change their resistance R over time or thermal cycling between room
temperature and our lowest accessible temperatures. We found that they
agreed with the original calibration within 1 mK. We also checked that their
calibration was negligibly aected by the magnetic eld
7
. Figure (3.3)(a)
shows R vs T for one of the thermometers used in our heat capacity studies.
(The shape is characteristic for low-T resistance thermometers). The data
were tted by a polynomial [open circles in Fig. (3.3)(b)] of the form
ln T =

i
C
i
(ln R)
i
, (3.1)
where C
i
are constants. (The method of least squares was utilized and the
same weight was given to each calibration point.) A plot of temperature
deviations T
th
= T
th
T vs T is shown in the inset to Fig. (3.3)(b). Here
T is the measured temperature and T
th
is the temperature calculated from
the observed resistance R and a 7th degree polynomial of the form given by
Eq. (3.1). The fractional deviation of the temperatures T
th
/T was below
2% for 20 mK T 1 K. In the event of unsatisfactory performance by
a carbon paint resistor, the carbon paint could be easily removed from the
sample upon immersion in its base solvent (isopropanol).
We now attempt to quantify the heat capacity of carbon paint resis-
tors. Following Ref. [110], the paint is dabbed using a thin copper wire,
in two spots of total mass 0.50 0.25 mg. Measurements of the specic
heat of graphite at low temperatures provide an experimental value about
6 10
7
J/(gK) at T = 400 mK, and show that well-ordered graphite does
not have a Schottky anomaly
8
. Our results seem to indicate that, in the in-
vestigated T-range, the heat capacity of the carbon paint resistors employed
in the present experiments is dominated by the electronic contribution. The
estimated heat capacity of the carbon paint is consequently expected to
range between 30 and 110 pJ/K at T 100 mK, consistent with the total
heat capacity signal. Based on estimations shown in Table (3.3), we con-
clude that in our experiments the heat capacity of carbon paint resistors is
negligibly small when the measured C C
QW
n
, but it dominates C when
the samples heat capacity reduces to (C
l
+C
e
).
7
While keeping the mixing chamber temperature constant, the temperature reading of
the carbon resistors showed < 3% monotonous deviation in the investigated B-range.
8
M.S. Dresselhaus, G. Dresselhaus, K. Sugihara, I.L. Spain, and H.A. Goldberg,
Graphite Fibers and Filaments, (Springer, Berlin, 1988), p. 106.
3.1. INTRODUCTION 103
Figure 3.3: Calibration data for carbon paint thermometers (see text).
104 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Heat leak
For sample M242, the thermometer was electrically connected with four
8 m-diameter NbTi wires. The electrical leads of the heater consisted of
one 60 m-diameter NbTi wire and one 8 m-diameter NbTi wire. These
NbTi wires (45% Nb by weight) also served as thermal link to the heat sink
and mechanical support; they remained superconducting in the investigated
B-range. Further, the sample M242 has been mechanically supported by a
peruoroalkoxy-insulated 250 m-diameter copper wire. For sample M280,
the thermometer and heater were electrically connected with graphite ber
bundles which also served as thermal link to the heat sink and mechanical
support. We have used ex-polyacrylonitrile (GY-70) graphite bers; they
have high Youngs modulus and their low-T thermal conductivity is small,
comparable with that of NbTi laments. Further, the sample M280 has
been mechanically supported by a small rubber O-ring shaped in the form
of a crown and a ne nylon tread [Fig. (3.2)(bottom)]. Various properties
of materials used for samples thermal anchoring are listed in Table (3.4).
The external thermal conductance could be easily adjusted by varying
the number of NbTi laments (or graphite bers) in order to obtain a con-
venient time constant for the estimated C in the T and B-range under
study. The thermal conductivity of graphite bers is obtained
9
from the
saturated low-T electrical resistivity by virtue of Wiedemann-Franz law:
= L
0
T, where L
0
=
2
k
2
B
/(3e
2
) = 2.45 10
8
(W/K
2
). The thermal
conductivity of NbTi wires was measured down to very low T and the fol-
lowing empirical law
10
was derived: 210
2
T
2
(W K
1
m
1
). Lower
bound estimates of the thermal conductance K
ext
of the actual heat leaks
in M280 and M242 sample calorimeters are given in Table (3.3).
In the discussion of the heat capacity of the thermal link, I want to put
aside as really relevant the fact that both NbTi wires and GY-70 graphite
bers display Schottky heat capacity anomalies. The carbon bers display
a Schottky heat capacity anomaly at low magnetic elds (B 0.1 T), prob-
ably due to the presence of magnetic impurities such as Fe. At the present
there is no explanation in terms of a microscopic model for the origin of
the hump in the heat capacity of graphite bers at low-B. The nuclear-spin
contribution of Nb atoms dominates the heat capacity of NbTi wires at high
9
V. Bayot, Conduction et Desordre dans le Graphite, Ph. D. Thesis, Universite
Catholique de Louvain, 1991, p.96.
10
J.R. Olson, Thermal Conductivity of Some Common Cryostat Materials between 0.05
and 2 K, Cryogenics 33, 729 (1993).
3.1. INTRODUCTION 105
magnetic elds (B 10 T). The estimated Schottky nuclear heat capacity
of Nb atoms at B = 10 T and T = 0.1 K in our M242 sample calorimeter is
1.3 10
5
pJ/K, which is comparable to C
QW
n
[Table (3.3)].
Table 3.4: Various properties of materials used for the thermal anchoring of
the sample in heat capacity experiments: mass density (kg m
3
), thermal
conductivity (W m
1
K
1
) at T = 0.1 K, specic heat (J kg
1
K
1
) at
T = 4.2 K, and electrical resistivity ( m) at T = 4.2 K.
Material Thermal
conductivity
Mass density Specic
heat
Resistivity
NbTi 0.2 10
3
6.2 10
3
0.87 0
GY-70 0.2 10
3
1.9 10
3
0.25 9 10
6
The simplied model of the calorimeter system
To keep the model for the calorimetric system at a manageable level of com-
plexity, it is necessary to make few simplifying assumptions: (1) the heat
leak has nite thermal conductance and zero heat capacity
11
, (2) the ther-
mometer and the heater are gifted with ethereal qualities, i.e., they are in
perfect thermal contact with the sample and have zero heat capacities
12
, (3)

int
is negligible when the samples heat capacity is dominated by the lattice
contribution (alternatively,
int
is very long when samples heat capacity is
dominated by the nuclear-spin contribution), (4) there is no heat transfer be-
tween the sample and its surroundings by radiation, (5) C(T +T) = C(T)
and (T + T) = (T) for small T
13
, and (6) at any T and B, electrons
are in thermal equilibrium with the lattice, i.e., the electron spin-lattice
relaxation time is the shortest time scale of the system.
11
The approximation of zero heat capacity for the heat leak is discussed in Sec. (3.1.4).
12
The heat capacity of carbon paint resistors is discussed in Sec. (3.1.4).
13
In all our calorimetric experiments, the temperature dierences are below 10% of the
mean temperature. At a given T and B, various nite heat capacities and thermal con-
ductances are considered as being constant within the periods of heating and T-response.
106 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
3.2 Calorimetry in the integer QHE regime
Before exploring the domain of lling factors close to unity, we stop along
the integer QHE to survey heat capacity measurements performed in this
regime. We have seen in Chapter 1 that T(E
F
) of a 2DES in a perpen-
dicular, quantizing magnetic eld exhibits 1/B-periodic oscillations due to
the formation of disorder-broadened Landau levels (LLs); T(E
F
) is max-
imum at half-integral and it is minimum at integral . The oscillating
T(E
F
) induces oscillations in many physical properties of the 2DES such
as electrical resistivity, magnetization, thermal conductivity, and specic
heat. Thermodynamic quantities (such as heat capacity) probe the total
DOS, as for thermal equilibrium properties the dierence between the lo-
calized and extended electronic states is of no importance. Quantitative
analysis of various measurements [30, 112] and, in particular, of heat capac-
ity experiments [42, 110], paint an extremely complex picture for the DOS.
Specically, the B-dependence of the LL width, the functional form of the
LLs, and the number of localised states between two adjacent LLs remain
largely unsettled.
We successfully carried out heat-capacity measurements in the integer
QHE regime at dilution refrigerator temperatures using the ac calorimet-
ric method [9]. Our results are qualitatively consistent with heat capac-
ity data reported early on [42], suggesting that the total DOS consists of
Gaussian peaks at the unperturbed LL positions superposed on a constant
background. It should be noted that information about the DOS in sam-
ples M242 and M280 could be also extracted from existing magnetization
and thermal conductivity data [8]. Since it is of notoriety that dierent
techniques yield dierent results for DOS, one expects that the comparison
of results of various techniques on the same samples will provide a consis-
tency test for quantitative determinations of the DOS. Why this thesis will
skip over such a quantitative data analysis? Firstly, uttermost care must be
taken in comparing the observed magneto-oscillations, as dierent aspects
of the DOS dominate the line shape of dierent thermodynamic quantities.
For example, the maxima of oscillations appear in magnetization (or lattice
thermal conductivity) when E
F
is between the LLs, whereas they appear
in the heat capacity when E
F
is at the center of the LLs. The LLs widths
estimated from the magneto-oscillations of heat capacity and magnetiza-
tion/thermal conductivity are therefore not directly comparable. Secondly,
as a complete synopsis of the T-dependence of the electrical resistivity in
the integer QHE plateaus is lacking at the time of the writing, it was not
possible to learn much from magnetotransport about the DOS between LLs.
3.2. CALORIMETRY IN THE INTEGER QHE REGIME 107
3.2.1 The steady-state, ac-temperature calorimetric method
In the ac calorimetry [98] the heater is driven by an ac voltage v
H
=
V
H

2 cos (
H
t/2), where V
H
stands for the rms amplitude of the cos-
inusoidal voltage. The corresponding heat generated in the sample is
P(t) = P + P cos (
H
t), where P = V
2
H
/R
H
is the peak power and R
H
is the heaters resistance. This heating signal raises the temperature of the
sample above T
s
, where it oscillates about a steady-state oset temperature.
The temperature dierence between the sample and the heat sink could be
written as T(t) = T
dc
+ T
ac

2 cos (
H
t ), where T
dc
is the con-
stant oset in the samples temperature, T
ac
is the rms amplitude of the
resulting temperature oscillations, and is the phase shift with respect to
heating oscillations. Assuming that both heat and temperature variations
are small and linearizable
14
, the thermal equation for the system is
C

_
T(t)

t
= P(t) K
ext
T(t). (3.2)
In the steady state, the dc component of the heating is dissipated through the
thermal link (T
dc
= P/K
ext
) and the ac component of the heating probes
the heat capacity. In the rst-order approximation, the heat capacity of the
sample is given by
C = P/(

2
H
T
ac
). (3.3)
Diagnostic procedure
The approximation that T
dc
is constant throughout the sample requires
that
ext
> 1/
H
>
int
. In physical terms this condition means that the
heat ux is cycled into the sample quickly enough that the thermal response
is measured before the heat dissipates to the exterior, but slowly enough
that the temperature response is uniform within the sample
15
. One always
measures the frequency prole of the calorimeter to determine an operating
frequency within a plateau region where neither
ext
nor
int
dominates the
thermal response. When the ac calorimeter has a plateau over at least one
decade in frequency together with sharp low-frequency and high-frequency
rollos, a good choice for the measuring angular frequency is (
int

ext
)

1
2
.
14
In practice, the constant rise in temperature corresponds to a drop in the ther-
mometers resistance. The corresponding drop in the thermometers resistance at
H
was checked to be proportional to V
2
H
at
H
/2.
15
The phase shift is close to /2 for this intermediate frequency range, it is essentially
zero for small
H
, and it approaches /4 for large
H
.
108 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
But when the plateau is only a broadly rounded hump and the time con-
stants of the calorimeter cannot be easily identied, the measuring frequency
is guessed according to few practical rules. It should be as high as possible
to avoid 1/f noise in the electronics and low-frequency limitations of the
lock-in detector. The lower limit of the measuring frequency is indicated by
the onset of deviations from the linear behavior describing the calorimeter
response at low frequency [110]. The upper limit on frequency is dictated
by the need for thermal equilibrium in the sample. The thermal relaxation
length

= (2D

/
H
)
1/2
should be greater than the largest dimension of
the sample.
Figure 3.4: Calorimeter response f
H
T
ac
(or 1/C) vs f
H
=
H
/(2) for
samples M280 (a) and M242 (b). The arrows indicate the measuring f
H
and
the dotted lines show the linear thermal response at low frequencies [110].
3.2. CALORIMETRY IN THE INTEGER QHE REGIME 109
In practice, for any set of external parameters (T and B), we optimize the
frequency for heat capacity measurements by scanning f
H
T
ac
or, equiv-
alently, 1/C as a function of f
H
=
H
/(2). In Fig. (3.4), the product
between f
H
and T
ac
[panel (a), sample M280] and 1/C [panel (b), sample
M242] are plotted vs f
H
to determine the bounded range in the frequency
domain where C is given by Eq. (3.3). The similar thermal response observed
at low frequency is not unexpected as the estimated K
ext
and C are nearly
identical for samples M242 and M280. In our experiments, plateau signa-
tures are observed around f
H
20 Hz, which was chosen as the measuring
frequency
16
. Fortunately, in the integer QHE regime, over the investigated
T- and B-range, the changes in heat capacity are small and a strong signal
could be maintained with only a single operating frequency. In the B-range
near = 1, however, C increases dramatically, and the plateau is expected
at much lower frequencies. As the lower limit of the frequency range of the
lock-in ampliers (PAR 124A) used in this study is 2 Hz, no plateau signa-
tures were found in a scan of f
H
T
ac
vs f
H
near = 1 and a heat-pulse
technique (see below) was used for measuring heat capacity.
3.2.2 Results and discussion
The heat capacity data presented in Fig. (3.5) demonstrate the oscillatory
behavior of C
e
for a 2DES in the integer QHE regime. We note that heat
capacity oscillates with the correct phase; i.e., it exhibits minima at integer
. The average density of the 2DESs, inferred from the position of the
minima, is consistent with the magneto-transport data. The uncertainty
in the background subtraction usually precludes the in-depth analysis of
the 1/B oscillations; only the amplitudes of magneto-oscillations are well
determined and can be used to gain insight into the DOS. Even though
C is dominated by the lattice and addenda contributions, up to 10%
oscillations - coming from the 2DES, are clearly observed. It should be
noted that the linear T-dependence observed at half integers in Fig. (3.5),
namely, the scaling of the measured C with the temperature, corresponds
to the usual Fermi-liquid-like behavior arising from intra-Landau-level con-
16
As discussed above,
int
can be estimated from the thermal diusivity data:
int

L
2

/D

. The phonon mean free path

of bulk GaAs depends on the doping. For n-


type GaAs,

is 300

A at room temperature and it rises when T is lowered. For the
highest doping it saturates at low-T about 0.5 10
3
m, whereas for the lowest doping
it saturates at low-T about 10
3
m. Henceforth, taking v

3000 m/s, for sample M280


(

10
2
m) the estimated
int
ranges between 10
3
s and 2 10
2
s. The latter
value is very close to
ext
, determined from the empirical analysis of the thermal signal,
explaining why in our experiments the theoretical plateau is conspicuously lacking.
110 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.5: [(a) and (b)] Heat capacity C vs B in the integer QHE regime.
3.2. CALORIMETRY IN THE INTEGER QHE REGIME 111
tribution to C
e
when E
F
lies within a broadened Landau level. Because the
C line shapes are similar for materials M242 and M280, we mainly discuss
below the results for sample M242 [Fig. (3.5)(a)] and, in particular, those
obtained at T = 92 mK [Fig. (3.6)].
Figure 3.6: Line shape of C
e
in the integer QHE regime after the subtraction
of the background as a Schottky term (sample M242 at T = 92 mK).
In order to further shine light on the experimental ndings, we shall use a
Gaussian model of the DOS [Eq. (1.16)], in which, neglecting the overlapping
between LLs, the DOS peaks are given by [eB

/(h
G
)][1/(

2)]. Thus, we
obtain

G
[eV] = 1.5 10
4
n
2D
C
e
[pJ/K]
A[cm
2
] B[T] T[K], (3.4)
where A is the area of the sample and n
2D
= 100 is the number of 2D
electron layers. The results for
G
are given in Table (3.5). For sample
M242 at T = 92 mK and = 5/2 the amplitude of heat capacity oscillations
is C
e
= 3.9 pJ/K, implying that the LL rms width is
G
( = 5/2) =
0.4 meV. Similarly, at T = 92 mK and = 7/2 ( = 9/2), where C
e
=
2.6 pJ/K (C
e
= 0.8 pJ/K), we infer
G
( = 7/2) = 0.45 meV [
G
( =
9/2) = 1.1 meV]. Strictly speaking, an amplitude of oscillation measures
112 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
the dierence between C
e
with E
F
at the peak and with E
F
between peaks.
If the DOS of the 2DESs, and therefore C
e
, is not negligible between LLs,
so that peak amplitudes are oset, then
G
obtained here is only an upper
bound
17
.
Table 3.5: The measured C
e
(pJ/K) at half-integer lling factors are given
at various temperatures for samples M242 and M280. The uncertainty in
C
e
is less than 0.2 pJ/K. For sample M280 note that, mainly due to the
strong B-dependence of the background [Fig. (3.5)(b)], there is no percep-
tible dierence between C
e
measured at 68 and 50 mK. Estimates of
G
(meV) [Eq. (3.4)] are also included.
= 5/2 = 7/2 = 9/2
C
e

G
C
e

G
C
e

G
M242 (92 mK) 3.9 0.40 2.6 0.45 0.8 1.1
M242 (62 mK) 2.9 0.35 2.2 0.35 - -
M280 (68 mK) 5.2 0.19 2.8 0.25 - -
M280 (50 mK) 5.2 0.14 2.8 0.18 - -
Comparison with previous heat capacity data [110] gives a LL width at
= 5/2 (and 7/2) about ve times smaller for our sample. This is consistent
with the lower disorder in our sample as evidenced by the presence of minima
in C
e
at odd down to = 5. Furthermore, the lower values of
G
in sample
M280 than those determined in sample M242 [Table (3.5)], correspond to
narrower, better-dened Landau levels, indicating an improved quality of
the 2DES in sample M280. It is worth noting that the disorder broadening

0
computed from the electrical resistivity data at B = 0 is in sharp conict
with the inferred Gaussian LL width at = 5/2 and 7/2. (
0
is one order
of the magnitude smaller than
G
[26].)
17
We recall that the short-range scattering theory [3], predicts that the LL width follows
the law
SR
(B) = [2
c
/(
0
)]
1/2
, i.e., at any B the LL width is uniquely described by

0
. For sample M242,
SR
(B) [meV] = 0.25 B
1/2
(with B in T). The calculated
SR
(B)
at = 5/2, 7/2, and 9/2 are smaller but comparable to the Gaussian LL widths obtained
from the amplitude of heat capacity oscillations [Table (3.5)].
3.2. CALORIMETRY IN THE INTEGER QHE REGIME 113
Qualitative analysis of the measured C
e
line shape also yields rough
estimates of the exchange-enhanced g factor [g(B)] at odd lling factors.
In Fig. (3.6), we see that C
e
has a weak minimum near = 5. The LL
width near = 5 must be comparable to the spin-splitting between LLs
E
Z
= g(B)
B
B, as the LLs are spin-resolved. Our upper bound of
G
( =
9/2) = 1.1 meV then implies g(B) 15.
In agreement with previous heat capacity experiments [109, 110, 111],
two important conclusions can be drawn from our data: (1) the zero-
magnetic eld mobility does not seem to parametrize the LL broadening very
eectively and (2) there is a g factor enhancement in the LL spin-splitting
of a 2DES. As in Refs. [42, 110], self-consistent numerical calculations and
multiple-parameter ts are necessary to determine C
e
, and thereby T(E
F
),
for arbitrary B. Our preliminary analysis (not shown here) suggests that
C
e
(B) could be reasonable well tted in the Gaussian model with
G
(B)
varying periodically in 1/B. In optimal ts to our data the peak heights in
the B-dependence of the DOS and C
e
are proportional to the inverse of the
Landau level width.
Finally, we wish to remark on the sudden and dramatic decrease of
the amplitude of C
e
oscillations in the N = 2 Landau level (4 6)
[Fig. (3.6)]. Since Shubnikov-de Haas oscillations of the longitudinal resis-
tance persist to much larger lling factors in the present samples, the in-
triguing behavior of C
e
suggests that some previously unanticipated physics
is at work. Recently, Koulakov, Fogler, and Shklovskii
18
have proposed
that in a clean 2DES in the N = 2 and higher LLs the uniform electron
liquid may be unstable against the formation of charge density waves. They
further suggest that near half lling of the LL the charge density wave is a
unidirectional strip phase having a wavelength of order of the cyclotron
radius. In this strip phase the electron density in the uppermost LL alter-
nates between zero and full lling. At = 5/2 this implies that there are
stripes of the incompressible states at = 4 and = 5. It is plausible to
consider that, if the stripes are coherent over the macroscopic size of our
samples, they provide a intrinsic source of inhomogeneity which smooths
away heat capacity oscillations in the higher LLs (N 2).
18
A.A. Koulakov, M.M. Fogler, and B.I. Shklovskii, Charge Density Wave in Two-
Dimensional Electron Liquid in Weak Magnetic Field, Phys. Rev. Lett. 76, 499 (1996).
114 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
3.3 The holy nuclear heat capacity of GaAs
QWs
The interest in heat capacity magneto-oscillations has been superseded by
the heat capacity data around = 1. At low temperatures (T 300 mK)
and in the B-range near = 1, C unexpectedly increased by orders of
magnitude compared to the low-B data. The observation of a giant low-T
heat capacity shifts our attention to the situation when the NSS contributes
to the measured heat capacity. We shall see that, indeed, our heat capacity
observations in the vicinity of = 1 are quantitatively described by a simple
model which takes into account the coupling between the NSS and the lattice.
Further consequences of the nuclear spin-lattice coupling near = 1 will
reward us with fundamental knowledge on the low-lying excitations of the
2DES.
3.3.1 The time-constant calorimetric method
To facilitate these developments, it is useful now to establish the method-
ology for the analysis of the heat capacity experiments in the practically
important case when the QWs nuclei contribute to the measured C. The
essential key to our time-constant calorimetric technique is that the thermal
conductance associated to the nuclear spin-lattice relaxation process is given
by K
QW
n
= C
QW
n
/T
1
. Here C
QW
n
is the Schottky heat capacity of the QWs
nuclei that couple to the lattice and T
1
is the nuclear spin-lattice relaxation
time which governs the heat transfer between the NSS of GaAs QWs and
the lattice.
In our thermal relaxation experiments, a constant power P is rst in-
jected in the lattice. After a period of time (t) much longer than
ext
, the
temperature of the nuclei that couple to the lattice (T
QW
n
) is equal to lat-
tices temperature T [t < 0 in Fig. (3.7)]. Consequently, we determine the
external heat leak as K
ext
= P/T
1
, where T
1
is the dierence between
the steady-state temperature of the lattice (T at t < 0) and the temperature
of the heat sink (T
s
). At t = 0 the power is turned o and the thermal relax-
ation of the nuclei then occurs through 1/K = 1/K
ext
+ 1/K
QW
n
according
to
C
QW
n

_
T
QW
n
(t)

t
= KT
QW
n
(t), (3.5)
3.3. THE HOLY NUCLEAR HEAT CAPACITY 115
Figure 3.7: The lattice temperature T vs time t in a relaxation heat capacity
experiment (sample M242). The two-parameter exponential t to the T
decay at t > 0 (full line) gives and T
2
. The inset shows the thermal circuit
of the sample with thermal conductances K
ext
and K
QW
n
. The constant
power injected in the lattice at t < 0, together with T
1
, determines K
ext
.
116 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
where T
QW
n
(t) = T
QW
n
(t) T
s
. Introducing the time constant =
C
QW
n
/K, the formula
T
QW
n
(t) T
s
= T
1
exp (t/), (3.6)
describes the evolution of T
QW
n
with t. Since in our case C
QW
n
is much larger
than any other contribution to the measured C, at t = 0 the temperature
drops from T
s
+ T
1
to T
s
+ T
2
, and for t > 0 the same amount of heat
ows through K
QW
n
and K
ext
K
_
T
QW
n
(t) T
s

= K
QW
n
_
T
QW
n
(t) T(t)

= K
ext
_
T(t) T
s

. (3.7)
Equations (3.6) and (3.7) give
T(t) T
s
=
K
K
ext
T
1
exp (t/), (3.8)
which can be used to analyze experimental data as illustrated in Fig. (3.7).
Finally, we obtain the heat capacity C C
QW
n
as
C =
P
T
1
_
T
2
T
1
_
, (3.9)
and the nuclear spin-lattice relaxation time as
T
1
=
_
1
T
2
T
1
_
. (3.10)
Equation (3.9) shows that a correction factor must be introduced to the
naively expected C = P/T
1
result as soon as a sharp temperature drop
is observed at the beginning of the thermal relaxation curve, which could
be still modeled by a single-exponential T-decay for t > 0. Equation (3.10)
tells us that T
1
is also propitiously measured in the above-described thermal
relaxation method. This model, usually described as the lumped
int
-
eect, has been discussed in the literature [6, 91] in the context of a poor
thermal contact between the sample and the substrate which has the heater
and the thermometer. To our knowledge, this is for the rst time when T
1
is
simply determined from the analysis of cooling curves exhibiting the lumped

int
-eect.
3.3. THE HOLY NUCLEAR HEAT CAPACITY 117
Limitations of the time-constant calorimetric method
The modied time-constant calorimetric method suers from the general
problems of thermal relaxation calorimetry
19
, and it is, in fact, applicable
only in very restrictive conditions
20
. First, attention has to be paid to
the asymptotic limit T
2
/T
1
0. In order to get a decent measure of
[6], one needs 0.1 < K/K
ext
, which corresponds to 0.1 < T
2
/T
1
. The
second limiting case, when T
2
/T
1
1 (K/K
ext
1), reduces to the
conventional case introduced in Sec. (3.1.1), in which, for a single dominant
relaxation time, one can retrieve only one parameter: the total heat capacity
C. Therefore, in order to reduce the inaccuracy in the determination of the
two right terms of Eqs. (3.9) and (3.10), we extract C and T
1
only in the
range 0.1 < T
2
/T
1
0.8. In practice, since K
QW
n
(and hence K) is
usually strongly T- and B-dependent, for a given K
ext
, thermal relaxation
curves of the form shown in Fig. (3.7) are observed only in a limited T-
and B-range. Clearly, one major diculty in doing this type of calorimetric
experiments is the suitable choice of the heat conductance of the thermal
link. We have checked that changing K
ext
by up to a factor of 5 has no
eect on the measured values of C and T
1
.
It may happen that, even though the nuclear-spin lattice coupling does
not provoke a lumped
int
-eect, it leads to a pronounced non-exponential
temperature decay for t 0 (the so-called distributed
int
-eect). This
problem has been solved by a number of authors [4, 6]; the relaxation curve
is usually approximated as a sum of two exponentials. Consequently, the
analysis is more involved. Andraka et al. [4] discussed in detail the nuclear-
spin lattice coupling in relaxation calorimetry from the point of view of the
distributed
int
-eect. In order to obtain the quantities of interest (C, C
QW
n
,
and T
1
), they propose a general method to analyze thermal relaxations with
two dominant time-constants
21
. Occasionally, we have observed cooling
curves displaying an abnormally high initial slope compared to the rest of
the T-decay, indicative of the distributed
int
-eect. In this case, we simply
obtain an estimate of C from the single (largest) time constant as C =
P/T
1
.
19
There is one basic limitation of the time-constant calorimetric method. Near phase
transitions, the simple exponential T-decay becomes complicated. In this case, not only
T(t) must be measured but also the time derivative

T(t) must be known to obtain C.
20
We note that in our experiments, accurate determinations of time constants below
10 s are unfeasible.
21
These authors scrupulously take into account, in addition to the nuclear-spin heat
capacity, the nite heat capacity of the lattice C
l
in their model for the heat ow. For
C
l
= 0, rearrangement of Eqs. 5 and 6 in Ref. [4] leads to Eq. (3.7).
118 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.8: C vs B at = 0

for T = 60 mK () and T = 100 mK (),


showing orders of magnitude enhancement near = 1 over the low-B data.
3.3. THE HOLY NUCLEAR HEAT CAPACITY 119
3.3.2 Results
As a function of increasing B, in addition to oscillations associated with
the 2DESs oscillating DOS at the Fermi level [Fig. (3.5)], a giant low-T
heat capacity is observed around = 1 [Fig. (3.8)]. Relative to its low-B
magnitude ( 10 pJ/K), C exhibits up to 10
5
-fold enhancement near
= 1. As shown in Fig. (3.8), the = 0

data are qualitatively similar


for both investigated samples. A key point in the structure of Fig. (3.8)
is that C exhibits broad maxima at 0.8 and 1.2, and decreases
rapidly for 4/3 and 2/3. The fact that C maxima appear nearly at
the same in both samples is a prima facie evidence of two-dimensionality
playing an important role for the anomalous C behavior. Figure (3.8) also
captures the evolution of C vs B with temperature in samples M242 and
M280. In contrast to the low-B data where C decreases with decreasing T,
in the vicinity of = 1, C increases with decreasing T. For this particular
range of temperatures (T 60 mK), a noteworthy feature of heat capacity
data is that C(T) T
2
at maxima [see Sec. (3.4)].
The results presented so far suggest that the presence of the giant
heat capacity may well be a symptom of anomalous behavior of the 2DESs
near = 1. To verify this conjecture we repeated the experiments in tilted
magnetic elds. Figure (3.9) displays the evolution of C vs B

with the tilt


angle in samples M242 and M280. For each sample, heat capacity maxima
are seen at the same B

, making the 2DESs a necessity in explaining the


anomalous C behavior.
3.3.3 Discussion
Both the very large magnitude of C and the T
2
dependence point to the
nuclear Schottky eect which results from the entropy reduction of the NSS
with decreasing T when the thermal energy k
B
T is much larger than the
nuclear-spin energy level spacing
n
(see Appendix A). The observation of
the Schottky eect requires good coupling of the NSS to the lattice in order
to reach thermal equilibrium in the time scale of the experiment. Ordinarily,
this coupling is provided by electron spin-ip excitations and further relax-
ation to the lattice. Consequently, while the eect is commonly observed in
metals, it usually remains undetected in high purity materials with low free-
carrier density [22]. At rst sight, because of their high purity and the low
free-carrier density in the GaAs QWs, our multiple-QW heterostructures
should not be good candidates for the observation of a nuclear Schottky
eect. While this is supported by the low-B data [Fig. (3.5)] where only the
120 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.9: C vs B

at indicated tilt angles. (Top) Sample M242 and


T = 100 mK. (Bottom) Sample M280 and T = 60 mK. The dashed curves
represent the calculated nuclear heat capacity of GaAs QWs (see text) for
= 30

(M242) and = 46

(M280).
3.3. THE HOLY NUCLEAR HEAT CAPACITY 121
lattice and 2D electron layers contribute to the measured C, surprisingly a
Schottky behavior is observed near = 1. Why? This thesis proposes the
following answer: the contribution of GaAs QWs nuclei to the measured
C near = 1 is attributed to a Skyrmion-induced, strong coupling of the
NSS to the lattice. First and foremost, the key role of Skyrmions for the
observation of the Schottky behavior is supported by the absence of the
nuclear spin contribution to our measured C for 1.3 and 0.7. At
these lling factors Skyrmions are no longer present in the ground state of
the 2DES. Secondly, the angular dependence of heat capacity clearly shows
that the heat capacity anomaly relates to (rather than total B) and thus,
reects intrinsic properties of the 2DES.
We now turn to a more quantitative interpretation of the data. The
dashed curves in Fig. (3.9) clearly indicate that the calculated C
QW
n
(see
Appendix A) is semiquantitatively consistent with both the size and the
overall B
2
dependence of the experimental data. A useful, dimensionless
parameter in describing heat capacity data is the ratio of the measured C
to C
QW
n
. This ratio provides a measure of the QWs nuclei which strongly
couple to the lattice, and hence signals the presence of low-energy spin
excitations in the 2DES associated to Skyrmions. This manoeuvre allows
us to jettison C vs B graphs [Fig. (3.9)(bottom)], and to replace them by
vs plots [Fig. (3.10)]. Presented this way, the = 0

heat capacity data


are highly symmetric around = 1. We see in Fig. (3.10) that shows
maxima of the order of unity at 0.8 and 1.2. While the decrease in
very near = 1 is attributed to the decreasing density of Skyrmions [14],
its decrease very far from = 1, i.e., 1.3 and 0.7, is related to
the 2DES approaching llings where the Skyrmions are no longer relevant.
The non-vanishing at = 1 could arise from density inhomogeneities
across the measured specimens, resulting in a nite density of Skyrmions at
nominal lling factor = 1. More precisely, we have found that up to 70%
of the QWs nuclei may contribute to the measured C at = 1 in the time
scale of the experiment. It should be mentioned that for sample M280, at
maxima reaches values above one [Fig. (3.10)], implying that the measured
heat capacity exceeds that of the QWs nuclei. Besides the experimental
accuracy (10%) and uncertainty in well-width (10%), the barriers nuclei
may enhance the measured C because of the penetration of the electron wave
function into the Al
0.1
Ga
0.9
As barriers and because of nuclear spin diusion
across the QW-barrier interface in the time scale of the experiment.
122 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.10: = C/C
QW
n
vs at T = 60 mK and indicated tilt angles (sam-
ple M280). The -envelope for = 0

(dotted line), is shown for comparison.


3.3.4 Disappearance of Skyrmions
We now address the important theoretical prediction that a transition from
Skyrmions to single spin-ip excitations is expected above a critical g
c
[95],
and present heat capacity measurements performed as g was tuned by tilting
the sample (M280) in the magnetic eld [77]. These measurements reveal
the absence of the nuclear spin contribution of GaAs QWs to the measured
heat capacity as g exceeds a critical value g
c
0.04. This absence suggests
the suppression of the Skyrmion-mediated coupling between the lattice and
the nuclear spins as the spin excitations of the 2DES make a transition from
Skyrmions to single spin-ips above g
c
. Figures (3.10) and (3.11) capture
the evolution of vs with tilt angle at T = 60 mK. As seen in Fig. (3.10),
3.3. THE HOLY NUCLEAR HEAT CAPACITY 123
vs at = 46

, is nearly identical to the = 0

data
22
. On the other
hand, at = 71

, the data show a signicant asymmetry with respect to


the = 1 position and a broadening of the > 1 peak. For < 1, at
0.9 is reduced by a factor of 2 when compared to the = 0

value and
it vanishes for 0.8. Most remarkable, however, is that the magnitude
of at the > 1 peak is comparable to the = 0

data, implying a still


strong coupling of the nuclei to the lattice. This is a particularly noteworthy
observation as it highlights that the heat capacity is a very sensitive probe
of the low-energy spin excitations, and therefore Skyrmions. We recall that
this is a regime where both calculations and transport data reveal a small
Skyrmion size (K 3) and a very weak dependence of
1
on g [Fig. (2.11)].
When is further increased above 71

only by few degrees [Fig. (3.11)],


the nuclear heat capacity decreases dramatically for all investigated . For
74

, the nuclear heat capacity is no longer detectable up to the highest


studied tilt angle (77

). To bring into focus the evolution of the coupling


between the NSS of GaAs QWs and the lattice with (and g), we plot at
> 1 and < 1 maxima vs g [Fig. (3.12)]. The coupling due to low-energy
electron spin excitations is progressively suppressed for g 0.035 and it van-
ishes in the range 0.037 g 0.043. We believe that this behavior provides
evidence for the transition from Skyrmions to single spin-ip excitations at
g
c
0.04 in our sample. This transition is quite abrupt: the disappearance
of the nuclear spin contribution of GaAs QWs to the measured heat capacity
occurs in a narrow g-range. The measured g
c
is smaller than the theoretical
g
c
= 0.054 calculated for the Skyrmion to single spin-ip transition for an
ideal 2DES [23]. We recount that several factors, however, are expected to
reduce g
c
for a real 2DES, including the nite thickness of the electron layer,
LL mixing, and disorder. Indeed, calculations by Cooper [see Secs. (1.6.1)
and (2.2.3)] reveal that taking into account the nite z-extent of the 2DES
alone leads to g
c
= 0.047, closer to our experimental value. The inclusion of
the LL mixing will further push down g
c
to 0.042 [76]. After considering
the role of disorder, we may conclude that the agreement between theory
and experiment is remarkable. It is worth emphasizing that calculations
are usually performed at = 1 while the heat capacity data of Figs. (3.10)
and (3.11) provide values for g
c
in the full -range (0.8 1.2) where
Skyrmions anti-Skyrmions are expected in the ground state of the 2DES.
In particular, we observe that g
c
depends on ; it increases from g
c
0.037
at = 1.2 to g
c
0.043 at = 0.9. The dierent g
c
observed at < 1 and
22
We note that in sample M242, vs at = 53

(the highest investigated tilt angle)


is nearly identical to the = 0

data.
124 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.11: = C/C
QW
n
vs at T = 60 mK for 71

74

(sample
M280). The dashed lines are guides to the eye.
> 1 could be brought into connection with the absence of the particle-hole
symmetry around = 1, favored by the tilted magnetic elds. Finally, we
also recall the possibility of a LL crossing at high tilt angles in our sam-
ple. This LL crossing might be responsible for the asymmetry of the heat
capacity data with respect to the = 1 position at 72

.
3.3.5 Ramications
We report here an anomalous and unexpected behavior for the measured
heat capacity at intermediate tilt angles (50

66

) which shows ev-


idence for the complex dependence of Skyrmions on and g. As depicted
in Fig. (3.13), we observe a reduction of in a narrow -range compared
to the = 0

data. The vertical arrows in Fig. (3.13) point to the total


3.3. THE HOLY NUCLEAR HEAT CAPACITY 125
magnetic eld at which is most signicantly reduced. The anomaly moves
to higher with small increases in (from = 0.81 at = 50

to 1 at
= 66

), and shows an increasing intensity when compared to the = 0

data-envelope, e.g., at = 50

, at = 0.81 is reduced by a factor of 3,


while at = 66

, at 1 is essentially zero. The observed reduction


of shows that low-energy electron spin excitations are strongly aected
or even suppressed, which would signal the disappearance of Skyrmions for
limited - or B-ranges dependent on tilt angle. It is worth noting that the
heat capacity anomaly occurs for [ 1[ 0.2 and 0.02 g 0.03. This is
precisely the g- domain where, according to theory [Fig. (2.9)], the 2DES
attains dierent subtle phases near = 1, some of which do not support
Skyrmions. Even though the exact behavior of the anomaly might be spe-
cic to our heterostructure, it reveals the subtle inuence of on the spin
excitations of the 2DES near = 1 whose description will require further
theoretical and experimental work.
Figure 3.12: = C/C
QW
n
at > 1 () and < 1 () maxima is plotted vs
g = [g

[
B
B/[e
2
/(l
B
)] at = 0

, = 46

, and 71

(sample M280).
126 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.13: = C/C
QW
n
vs at T = 60 mK and indicated tilt angles (sam-
ple M280). The -envelope for = 0

(dotted line), is shown for comparison.


3.3. THE HOLY NUCLEAR HEAT CAPACITY 127
We now complete the discussion of the heat capacity results obtained in
the presence of an in-plane magnetic eld component by comparing them
with tilted magnetic eld magneto-transport data. Recently, Kronm uller
et al. [61] discovered that magnetotransport near = 2/3 is inuenced by
the nuclear polarization of nuclei inside the GaAs QWs through the Fermi
contact hyperne interaction [1]. The question arises if the nuclear spins are
also involved in the physics of the = 4/3 fractional QHE state, which is
the electronic conjugate of the = 2/3 fractional QHE state. Examining
the heat capacity data of Figs. (3.10), (3.11), and (3.13) we observe a
broadening of the > 1 peak, starting with 61

. Figure (3.11) shows


the reduction of the > 1 peak for 72

. Therefore, these data seem


to support the idea that a nuclear heat capacity is observed near = 4/3
for a nite range of tilt angles. Why? A possible escape from this enigma
is provided by the tilted magnetic eld magneto-transport data, exhibiting
a re-entrant behavior of the = 4/3 fractional QHE state. The evolution
of
4/3
with B is summarized in Fig. (3.14)(a), whereas the evolution of
with B near = 4/3 is presented in Fig. (3.14)(b). Our tilted-magnetic
eld heat capacity data near = 4/3 uncover a rapid nuclear spin-lattice
relaxation in the B-range where the = 4/3 fractional QHE state is absent.
This behavior signals the presence of nearly gapless spin uctuations in the
2DES when the = 4/3 fractional QHE state undergoes a spin phase
transition (i.e., from a spin-unpolarized ground state to a partially polarized
ground state). These nearly gapless spin uctuations permit then the strong
coupling of the 2DES to the GaAs QWs nuclei.
Under the given context of tilted magnetic eld magneto-transport and
heat capacity experiments near = 4/3, we take the opportunity to note
that Kronm uller et al. [61] observed that sweeping B upwards and down-
wards produces a rapidly observable hysteresis in the longitudinal resistance
near = 2/3. While it is now widely accepted that the hysteresis in the
magnetoresistance near fractional QHE states is provoked by transitions be-
tween dierent spin congurations of the 2DES, several concerns (such as
the role of the nuclear spins) are not clearly established [20, 33]. For ex-
ample, the nuclear spins may be responsible for the slow relaxation of the
magnetoresistance in the vicinity of various fractional lling factors [33].
Further research is necessary before we can decide that the onset of a van-
ishing excitation gap at = 4/3 in our samples suces for the observation of
both nuclear heat capacity and hysteretic behavior. Searching for T- and -
dependent hysteretic behavior near = 4/3 in samples M242 and M280 may
shed light on the participation of the nuclear spins in spin phase transitions
of unpolarized fractional QHE states.
128 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.14: (a)
4/3
vs B (sample M280). Panel (b) displays the B-
dependence of = C/C
QW
n
near = 4/3 at T = 60 mK. The grey regions
indicate the B-range where the = 4/3 fractional QHE is absent. The
corresponding tilt angles are given on the top axis.
3.4. SKYRMION LATTICES 129
3.4 Skyrmion lattices
3.4.1 Heat capacity measurements at very low temperatures
At very low temperatures (T 60 mK) and in a narrow -range ( 0.8
and 1.2), C is very large and the observed external time constant
ext
reaches extremely large values which exceed 10
4
s. Therefore, in the time
scale of our experiments, the sample is in the quasi-adiabatic regime
23
.
Figure 3.15: C vs T at B = 7 T ( = 0.81) is shown in the main gure in
a log-log plot (sample M242). The dashed line shows the T
2
dependence
expected for the Schottky model, and T
c
is marked by the vertical arrows
(see text). The inset shows a linear plot of C vs T at B = 6.7 T ( = 0.85).
23
All data discussed in this section were taken on sample M242.
130 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure (3.15) shows the T-dependence of C at = 0

for = 0.81 and


= 0.85. We observe that the C T
2
behavior is followed only down
to 60 mK. For T 60 mK, C increases faster with decreasing T, and
exhibits a remarkably sharp peak at T
c
before decreasing at very low T.
Figure (3.16) displays the T-dependence of C at = 30

for two lling


factors in the neighborhood of = 0.8. Our tilted magnetic eld data
indicate that the peak in C vs T could be associated to the 2DES, as it is
observed in the same range of lling factors (i.e., around = 0.8 and 1.2).
Figure 3.16: C vs T at = 30

and indicated values of (sample M242).


The T
2
dependence expected for the Schottky eect is shown as a dashed
line, and T
c
is marked by the vertical arrows. The -dependence of T
c
at
= 0

() and 30

() is shown in the inset. The lines are guides to the eye.


3.4. SKYRMION LATTICES 131
A prominent feature of data taken at = 30

is that T
c
sensitively depends
on . The inset to Fig. (3.16) shows that the peak temperature T
c
is max-
imum at 0.8 and 1.2 and decreases as 1. The sharp peak in
C vs T is not observed in sample M280, which might signal that it occurs
at lower temperatures in this lower density heterostructure; the Schottky
T-dependence [C(T) T
2
] is followed down to T 30 mK. We note that
for sample M280 we found a marked non-Schottky behavior of the heat ca-
pacity at very low temperature at = 71

and = 1.22 [Fig. (3.17)], which


is reminiscent of the sharp peak in C vs T observed in sample M242.
Figure 3.17: C vs T at = 0

( = 0.88, lled circles) and = 71

( = 1.22,
open circles) is shown in a log-log plot (sample M280). The dashed line shows
the T
2
dependence expected for the Schottky model.
132 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
The deviation of C from the T
2
dependence at T
n
/k
B
and the shape
of the observed peak are clearly not consistent with the Schottky model
which predicts a smooth maximum in C at T
n
/(2k
B
) [ 2 mK in our
case, solid curve in Fig. (3.18)]. Instead, the shape and sharpness of this
peak are suggestive of a phase transition in the 2DES. It is tempting to as-
sociate the sharp peak in C vs T observed at low T with the crystallization
of Skyrmions. The fact that our observed T
c
decreases as 1 is consis-
tent with the decreasing Skyrmion density which should reduce the melting
temperature of the Skyrmion solid. However, the details of the Skyrmion
liquid-to-solid transition are largely unknown. While it is expected that a
Skyrmion liquid-to-solid transition would aect
n
only very weakly, how
it would aect the nuclear spin dynamics is unclear.
The question arises what is the physical mechanism that aects the NSS
and gives rise to the anomalous C-peak. The substantial enhancement of C
at low T could be interpreted as an indication that either (1) more nuclei
couple to the lattice, or (2) the entropy of the coupled NSS decreases faster
with decreasing T than what is expected from the Schottky model. Picture
(1) relies on a modied coupling of the NSS to the lattice manifesting as an
enhanced nuclear spin diusion so that a larger number of nuclei contribute
to the heat capacity near T
c
. Picture (2), on the other hand, relies on
a Skyrmion-induced nuclear spin polarization which reduces the entropy of
the NSS. This is reminiscent of the dynamic nuclear polarization of the NSS,
for example when nuclear spins interact with spin-polarized paramagnetic
impurities [1]. While in the Skyrmion-liquid state there is no preferential
orientation of the electron spins, the transition to a pinned Skyrmion solid
could possibly induce a local preferential orientation of the electron spins
which, in turn, would polarize the NSS and thereby reduce the entropy.
We have performed quasi-adiabatic thermal experiments [10] revealing
that the mechanism responsible for the peak in C vs T is a dramatic en-
hancement of nuclear spin diusion across the QW-barrier interface. While
only the nuclear heat capacity of the QWs atoms is observed away from T
c
,
an increasing number of nuclei in the barriers contribute to the measured
heat capacity when T T
c
.
3.4.2 The variable nuclear thermal coupling model
While at high temperatures (T 60 mK) the measured value and T-
dependence of C are consistent with the calculated Schottky nuclear heat
capacity of Ga and As atoms in the QWs [solid curve in Fig. (3.18)], at lower
T, C exceeds the calculated value by a factor of up to 10 at T
c
. The peak
3.4. SKYRMION LATTICES 133
value of C appears consistent with the Schottky nuclear heat capacity of the
heterostructure if the nuclei of the barriers are also included [dotted curve
in Fig. (3.18)]. This observation suggests that the peak in C vs T might
Figure 3.18: Measured C vs T at B = 8.5 T and = 30

(sample M242).
The curves represent the calculated Schottky nuclear heat capacity of the
one hundred GaAs QWs (full curve) and the 100-period GaAs/Al
0.3
Ga
0.7
As
heterostructure (dotted curve).
come from the contribution of barriers nuclear spins to the measured C.
Since the nuclear spin-lattice relaxation rate is extremely weak in the bar-
riers at such low T, such an interpretation necessarily implies a variable
thermal coupling between the QWs nuclei and the barriers nuclei. The
above observations can be cast into a more concrete model as schematically
shown in the inset to Fig. (3.19)(a). Here K
QW
n
and K
B
n
are variable thermal
134 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
conductances which depend on both and T. They are normally negligibly
small so that one measures only the heat capacity of the lattice and the
2DESs
24
. Near = 1, however, K
QW
n
becomes large and the QWs nuclei
dominate the measured C, while near = 1 and T = T
c
both K
QW
n
and K
B
n
become large so that C of the entire heterostructure is measured.
3.4.3 Results of quasi-adiabatic thermal experiments
We have critically tested the above conjectured model by performing quasi-
adiabatic thermal experiments which utilize another noteworthy feature of
the heat capacity data; as seen in Fig. (3.16) inset, T
c
sensitively depends
on . In the framework of our model, this strong dependence allow us to
tune K
B
n
by varying B in order to cross the region near T
c
where K
B
n
is
maximum. If the barrier nuclei are not in thermal equilibrium with the
rest of the sample before the crossing, then the lattice T should be greatly
aected as T
c
is approached and the barrier nuclei thermally couple to the
QWs nuclei and the lattice. This is the key to understanding our results.
In these experiments, we rst x B at 8.5 T ( = 0.77) where the peak in
C occurs at T
c
= 42 mK. Starting from T 60 mK, we cool down the cold
nger to base temperature ( 10 mK) and the lattice slowly cools down
to T 32 mK by waiting for about 50 hours. At this temperature, the
measured heat capacity is much smaller than C observed at T
c
, implying
that the coupling to the barriers (K
B
n
) is quite small. Then, in order to
further improve the adiabatic conditions, the weak heat ow that results
from the T-dierence between the lattice and the cold nger is reduced by
increasing the cold nger temperature to 32 mK. Under these conditions,
we observe that the measured T is stable within 0.2 mK over periods
longer than
ext
, meaning that the sample is in the quasi-adiabatic regime.
In a rst experiment, whose results are shown in Fig. (3.19)(a) by open
circles, we started from such a quasi-adiabatic condition and swept B from
8.5 to 7.7 T, at a rate of 0.01 T/min. We observe that T rises from 31.7 mK
to 36.2 mK. B was then swept back to 8.5 T at the same rate; this led
to a rise of T to 39.5 mK. While the increase in T with increasing B may
be explained by the adiabatic magnetization of the nuclei coupled to the
lattice, according to
T
n
B, (3.11)
the T rise observed when B is lowered from 8.5 to 7.7 T is puzzling.
24
The sharpness of the peak in C vs T implies that K
B
n
0 fast as T deviates from T
c
.
3.4. SKYRMION LATTICES 135
Figure 3.19: (a) T vs B during the rst () and the second () quasi-
adiabatic experiments. The full and the dotted lines are guides to the eye;
the arrows show the direction of evolution. The dashed curve shows the
B-dependence of T
c
. (Inset) Schematic description of the coupling between
dierent parts of the sample. (b) T and B vs time in the second experiment.
B-sweeps (gray areas) separate hold times at the end of which T is recorded
(1 to 7) and summarized in (a) together with the nal T at B = 8.5 T (8).
136 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
In order to better understand the results of the above experiment, we per-
formed a similar experiment, starting from T = 32.6 mK, but here we swept
B in steps of 0.2 T from 8.5 to 7.3 T, with a hold time of 3 hours between
steps. The evolution of T during this second experiment is presented in
Fig. (3.19)(b) and summarized in Fig. (3.19)(a) by close circles. We note
that in each step, except for steps 2 3 and 3 4, the lattice T rises
during the B-sweeps and then decays to approximately the same value as
the one at the end of the previous step. These T rises are due to eddying
currents heating of the 2DES and the lattice during the B-sweeps. The heat
is evidently slowly absorbed by the nuclei during and following the sweep via
internal relaxation. Because of the very large heat capacity of the nuclear
spins, this heating (due to eddy currents) does not lead to an appreciable
increase in the lattice T at the end of the step. The key feature of this
experiment, however, is that during the 2 3 and 3 4 steps, the lattice
T increases signicantly while the sample is in quasi-adiabatic conditions.
Note in particular that, during the 2 3 step, the lattice T rises even after
the B-sweep is completed. Indeed, it appears that near T
c
the lattice is
heated internally.
3.4.4 The crystal of Skyrmions and the nuclear spin diusion
We now detail the predictions of the model for these particular experiments.
We expect that during the initial cool-down at B = 8.5 T, when T decreases
below T
c
= 42 mK, T
n
in the barriers (T
B
n
) remains close to 42 mK and T
n
of the QWs (T
QW
n
) decreases down to 32 mK due to strong coupling
to the lattice. Next, decreasing B below 8.5 T (steps 1 7) in adiabatic
conditions has two distinct consequences: (1) T
B
n
and T
QW
n
are reduced
due to adiabatic demagnetization of the nuclei [Eq. (3.11)], and (2) the
coupling between the QWs and the barriers will increase dramatically when
the dashed (T
c
vs B) curve in Fig. (3.19)(a) is crossed. While the former is
essentially monotonic in B, the latter is not and implies that, near the dashed
curve in Fig. (3.19)(a), T
n
should equalize over the entire GaAs/Al
0.3
Ga
0.7
As
heterostructure. Since the barriers nuclei were initially warmer than the
QWs nuclei, we expect a rise in lattice T near this crossing, as observed
experimentally. Finally, we attribute the T rise during the nal (7 8)
step to the adiabatic magnetization of the nuclei [Eq. (3.11)] coupled to the
lattice.
Beyond the good qualitative description of the experiments in Fig. (3.19),
the above model appears to provide a reasonable quantitative account of the
data also. Nuclear demagnetization [Eq. (3.11)] reduces T
B
n
from 42 mK
3.4. SKYRMION LATTICES 137
to 36 mK when B is lowered from 8.5 to 7.3 T. Since the barriers are
about 10 times thicker than the QWs, the rise in T
QW
n
from steps 1 7 in
Fig. (3.19) reduces T
B
n
only weakly ( 0.5 mK). Therefore, T
B
n
and T
QW
n
should merge towards a nal value close to 36 mK. This prediction is in
good agreement with experiments that give T 37 mK at B = 7.3 T
25
. The
nal increase of T from 37 mK to 43 mK when B increases from 7.3 to
8.5 T is accounted for by the adiabatic magnetization of the heterostructures
nuclei [Eq. (3.11)]. Lastly, the fact that C(T
c
) reaches the value expected
for the entire heterostructure is consistent with our model.
Our data also shed light on the thermal conductances K
QW
n
and K
B
n
as
they are related to the internal time constant (
int
) observed in our experi-
ments. Figure (3.20) presents a typical heat pulse/relaxation trace obtained
during heat capacity experiments near T
c
. In the T-range (T 1.2 T
c
) where
C exhibits a maximum, we measured C using the quasi-adiabatic technique,
ensuring that the measured C is not aected by possible changes in the
internal time constant of the system. We recall that C = Q/T where
T is the temperature increase resulting from the applied heat Q = Pt,
after internal relaxation is completed. The relaxation follows an exponential
decay characterized by
int
. The inset to Fig. (3.20) shows that
int
exhibits
two dierent, but remarkably constant, values above and below 30 mK:
1600 s and 900 s, respectively. We note that this crossover temperature
corresponds to the T below which C is reduced back to a value close to the
nuclear heat capacity of the 100-QWs [full curve in Fig. (3.18)].
In order to explain these observations, we consider three possible mech-
anisms responsible for the observed
int
. Heat diusion in the nuclear spin
system of the barriers is governed by the one-dimensional diusion equation
T
B
n
t
= D
n

2
T
B
n
z
2
, (3.12)
where D
n
10
17
m
2
/s is the nuclear-spin diusion constant in GaAs
26
and the z-axis is along the growth direction. Therefore, it gives an internal
time constant

d

r
2
D
n
, (3.13)
where r is the distance over which diusion takes place
27
. Since each barrier
25
Complementary experiments allow us to estimate the T increase expected from the
Eddy-current heating during the B-sweeps. Our estimate is 1 mK for the total T rise
during steps 17 or 78.
26
In our estimate here, we use D
n
for GaAs since D
n
for Al
0.3
Ga
0.7
As is not known.
27
Note that the actual boundary conditions used to solve Eq. (3.12) may add a prefactor
138 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
is surrounded by two QWs, we take r = 925

A (half the barrier thickness)


and Eq. (3.13) gives
d
800 s. The two other mechanisms involved in the
heat transfer from the lattice to the barrier nuclei, i.e., nuclear spin-lattice
relaxation in the QWs and diusion across the QW-barrier interfaces, result
in two additional time constants: T
1
and
i
, respectively. When C C
QW
n
(T 30 mK), the diusion across the QW-barrier interface is negligible and

int
should be determined by T
1
only. On the other hand, when C exceeds
C
QW
n
(30 T 50 mK) at least a fraction of the barriers nuclei couple to
the QWs nuclei. This implies that
d
and
i
should increase
int
.
Figure 3.20: T vs time during a heat capacity experiment in the quasi-
adiabatic regime at B = 8.5 T and = 30

(sample M242). The heat pulse


is followed by a relaxation characterized by an exponential decay (full curve)
giving
int
1500 s. The inset shows
int
vs T at the same B and .
This is consistent with the rise in
int
observed above 30 mK [Fig. (3.20)
to the right term of Eq. (3.13) which gives only a rough estimation of
d
.
3.4. SKYRMION LATTICES 139
inset]. Moreover, we note that the increase of
int
above 30 mK ( 700 s)
is comparable to the estimated
d
800 s, implying that
i
is rather small
once the barrier nuclei do couple to those in the QWs
28
.
We now remark on the physical origin of the peak in C vs T. Optically
pumped nuclear magnetic resonance (OPNMR) experiments [7, 103] show
that nuclear spin diusion from the QWs into the barriers is extremely weak,
except when optical pumping broadens the NMR (Knight shift) peak of the
QWs which then overlaps with the NMR peak of the barriers. The spectral
overlap allows spin diusion which is driven by nuclear magnetic dipole-
dipole coupling. Therefore, the enhancement of the nuclear spin diusion
across the QW-barrier interface near T
c
could originate from a broadening
of the QWs NMR line. In the Skyrmion-liquid phase, motional averaging
of the local spin polarization of the 2DES produces a single Knight shift
peak (see for details Chapter 4 and Appendix B). On the other hand, the
absence of motional averaging in the Skyrmion-solid phase should induce
both positive and negative Knight shifts depending on the local spin polar-
ization of the 2DES. Since above and below the peak in C vs T the diusion
across the QW-barrier interface is very weak, this implies that in both the
liquid and the solid Skyrmion phases, the Knight shift peak(s) do not have
signicant overlap with the NMR peak of the barriers.
One possible explanation for the C vs T peak is that the critical slowing
down of the spin uctuations in the 2DES, associated with the Skyrmion
liquid-to-solid phase transition, could induce a broadening of the QWs NMR
peak
29
. This broadening would induce spectral overlap of the QWs and
barriers NMR peaks and hence allow spin-diusion across the QW-barrier
interface only near T
c
.
The fact that
int
is constant in the T-range where C > C
QW
n
near T
c
implies that the strength of the coupling to the barriers remains constant
while C varies strongly. Therefore, we deduce that the smaller value of C
on the sides of the peak in C vs T comes from a reduced fraction of the
sample in which the barriers nuclei couple to the QWs nuclei. The best
explanation for this is that in a magnetically ordered Skyrmion state, there
are regions around Skyrmions where the electronic magnetization is perpen-
dicular to the external magnetic eld and the Knight shift vanishes. The
nuclear spins outside the QWs can come into equilibrium with the nuclei in
28
It is worth remarking that the parts of the sample for which
d

ext
do not con-
tribute to the measured C in the T and B range we are discussing. Since
d
10
5
s for
r > 10
6
m [Eq. (3.13)], the GaAs substrate contributes negligibly to C.
29
The critical slowing down causes the electron motion time scale to pass through the
NMR time scale.
140 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
these regions. The size of these regions is expected to peak at the Skyrmion
solid-to-liquid phase transition due to the critical slowing down of spin uc-
tuations. Hence, specic heat measurements suddenly see the nuclei outside
the QWs when a Skyrmion lattice forms. The complexity of spin textures in
the 2DES, together with the fact that the phase transition is not observed
directly but through its eect on the nuclear spin-lattice relaxation, does not
allow a simple analysis of the data in the framework of the existing theories
that describe the melting and magnetic disordering of the Skyrmion lattice.
It is worth emphasizing that values of the Skyrmions moment of inertia
and stiness to relative rotations to neighboring Skyrmions for the multi-
Skyrmion system extracted from calculations [24] suggest that the system is
close to criticality. The constant thermal relaxation rate for T 50 mK is
in accord with the predicted T dependence of T
1
1
in the quantum critical
regime [44]. To sum up, our results provide a semiquantitative phenomeno-
logical description of the peak in C vs T that may originate from a Skyrmion
solid-to-liquid phase transition.
3.5 The nuclear spin-lattice relaxation rate
We now briey discuss the T
1
1
results of our calorimetric experiments
30
.
Figure (3.21)(a) summarizes the T- and -dependence of T
1
1
in sample
M242, measured at = 30

. Interestingly, T
1
1
decreases when T is lowered
and, at a given T, a common T
1
1
is observed for 0.7 1.3. That is, the
nuclear-spin lattice relaxation is as rapid at = 1 as on the either side of
= 1 down to very low temperatures, consistent with the observation of a
large nuclear spin heat capacity at = 1 (0.3 0.7). In Fig. (3.21)(b) we
show Arrhenius plots of T
1
1
for various lling factors, revealing an activated
T-dependence.
The situation is rather subtle and a concise presentation of OPNMR
T
1
1
results is appropriate at this juncture (see for details Chapter 4 and
Appendix B). OPNMR experiments at 1 and 1 show no T-
dependence for T
1
1
in the temperature range 1.5 K T 4.2 K. In
this T-range, the nuclear-spin lattice relaxation is rapid on either side of
= 1 (T
1
1
0.05 s
1
). On the other hand, at = 1, OPNMR measure-
ments found a slow nuclear-spin lattice relaxation (T
1
1
< 0.01 s
1
), which
strongly depends on the temperature [Table (B.1)]. This peculiar behavior
of T
1
1
(T, ) has been attributed to Skyrmion induced low-lying spin-ip
30
An in-depth analysis of T
1
1
calorimetric data awaits completion at the time of the
writing.
3.5. THE NUCLEAR SPIN-LATTICE RELAXATION RATE 141
excitations in the 2DES [7, 103]. Clearly, this assignment is fully consistent
with existing calculations [24].
Our low-T T
1
1
calorimetric data open new vistas on the many-Skyrmion
ground state at T 0. One of the central queries here is to understand the
striking -dependence of T
1
1
near = 1, i.e., the conspicuous absence of
the slowing down of the nuclear-spin lattice relaxation at = 1. We believe
that, due to a nite density of Skyrmions at nominal lling factor = 1,
T
1
1
describes the average response of dierent spin-textured domains (with
the shortest time scales for nuclear spin-lattice relaxation) in the experi-
mental time window. It is simple to imagine that, under the exceptional
circumstance of Skyrmions present in the electronic ground state at = 1
and of an inhomogeneous
31
2DES, T
1
1
at nominal lling factor = 1 could
be constructed from averaging over the T
1
1
contributions at = 1 .
Recounting that T
1
1
varies linearly with [1 [ [24], it is easy to see that
the -averaged T
1
1
will have essentially no dependence on the lling factor
around = 1 and a fast nuclear spin-lattice relaxation is expected even for
T 0.
We now attempt to interpret the T-dependence of T
1
1
[Fig. (3.21)]
32
.
At low-T (T 60 mK), the experimentally determined T
1
1
never becomes
smaller than 1 10
3
s
1
for all investigated lling factors. This sat-
uration value is close to the T
1
1
observed near the integer lling factors,
in electron spin resonance experiments [28]. At higher temperatures, for
lling factors where Skyrmions are the expected ground state of the 2DES,
T
1
1
shows an activated T-dependence with a very small activation energy
0.3 K, consistent with a phonon-assisted mechanism through very low
energy spin-ip excitations in the 2DES [57, 105]. Our speculative inter-
pretation for this behavior is the following. The fact that Skyrmions might
be in a magnetically ordered liquid state at low temperatures [24], together
with disorder pinning, introduces a gap (
RR
) in the energy spectrum of
the 2DES. The essential physics will still be that the spin uctuations have
strong spectral density at frequencies far below the Zeeman gap [39]. The
presence of
RR
, in turn, induces an activated T-dependence for T
1
1
when
T
RR
[28, 105]. At high-T (T 0.3 K), the experimentally determined
T
1
1
saturates at 0.1 s
1
. We shall see in the next chapter that T
1
1
s
determined near = 1 in calorimetric experiments for T 0.3 K are
31
The inhomogeneity could simply be due to a non-uniform Skyrmion density.
32
We neglect here the step in the thermal relaxation rate at T = 30 mK [inset to
Fig. (3.20)].
142 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
Figure 3.21: (a) T
1
1
vs B (bottom axis) and (top axis) at = 30

and indicated temperatures (sample M242). (b) Arrhenius plots of T


1
1
at
= 1.1 () and = 0.77 (), revealing an activation energy
RR
= 0.32 K.
3.6. CONCLUSION 143
approximately equal to T
1
1
s observed in our standard NMR studies at
T = 1.5 K.
3.6 Conclusion
In this chapter we described heat capacity measurements down to very low
temperature (T 20 mK) and Landau level lling factor ( 0.5). As a
function of increasing magnetic eld, B, in addition to oscillations associated
with the 2DESs oscillating density of states at the Fermi level, we observe
a dramatic increase of the low-T heat capacity in the range 0.5 1.5.
For 0.8 and 1.2, C exhibits a striking T-dependence, including a
remarkably sharp peak at very low temperature (T
c
) suggestive of a phase
transition in the 2DES. We interpret these unexpected observations in terms
of the Schottky model for the nuclear-spin heat capacity of Ga and As
atoms which couple to the lattice via the 2DESs low-energy spin-texture
excitations (Skyrmions).
Heat capacity measurements on sample M280 uncover the disappearance
of the nuclear spin contribution to the heat capacity as the ratio g between
the Zeeman and Coulomb energies exceeds a critical value g
c
0.04. This
disappearance suggests the vanishing of the Skyrmion-mediated coupling be-
tween the lattice and the nuclear spins as the spin excitations of the 2DES
make a transition from Skyrmions to single spin-ips above g
c
. This g
c
is
somewhat smaller than g
c
= 0.054 expected from theoretical calculations for
an ideal 2DES. The discrepancy likely comes from corrections to g
c
because
of nite layer-thickness, Landau level mixing, and disorder in a real sam-
ple. Furthermore, our tilted magnetic eld heat capacity experiments reveal
the subtle and critical inuence of the in-plane magnetic eld component
on the ground and excited states of 2DESs near = 1 in GaAs/AlGaAs
heterostructures.
We discussed the physical origin of the sharp peak in C vs T in terms
of the possible Skyrmion solid-to-liquid phase transition. The sharp peak in
C vs T near = 1 can not possibly be due to the tiny amount of entropy
change in the Skyrmion lattice itself. Rather it is due to the nuclei in the
thick barriers between the GaAs quantum wells. Quasi-adiabatic thermal
measurements on sample M242 show that the state of the conned 2D elec-
trons profoundly aects the nuclear-spin diusion near Landau level lling
factor = 1. Our experiments provide quantitative evidence that a dra-
matic enhancement of nuclear-spin diusion from the GaAs quantum wells
into the AlGaAs barriers is indeed responsible for the sharp peak in the tem-
144 CHAPTER 3. HEAT CAPACITY EVIDENCE FOR SKYRMIONS
perature dependence of heat capacity near = 1. While only the nuclear
heat capacity of the QWs atoms is observed away from T
c
, an increasing
number of nuclei in the barriers contribute to the measured heat capacity
when T T
c
.
Chapter 4
Skyrmions Probed by NMR
Fit i treji, privegheat i. Potrivnicul vostru, diavolul, umbla,
racnind ca un leu, cautand pe cine sa nghita, caruia stat i
mpotriva, tari n credint a, stiind ca aceleasi suferint e ndura
si frat ii vostri n lume.
Sf. Apostol Petru
4.1 Overview
In recent years, the development of novel low-temperature spectroscopic
methods has made available a new set of tools for the study of spin-related
phenomena in the quantum Hall eect (QHE) regime. A principal task of
these spectroscopic experiments [2, 7, 62, 78] is the direct measurement of the
spin polarization T of two-dimensional (2D) electrons. In nuclear magnetic
resonance (NMR), T is determined from the (Knight) shift of NMR lines
which is a direct measure of the additional eective magnetic eld created by
the electronic spins at the position of the nucleus. Unfortunately, the NMR
signal from a conventional single quantum well (QW), containing a two-
dimensional electron system (2DES), is typically too weak to be observed.
145
146 CHAPTER 4. SKYRMIONS PROBED BY NMR
The reason for this is that the NMR signal is proportional to the product of
the number of nuclear spins and their average polarization; the number of
nuclei located in a single QW, which interact with 2D electrons, is small and
their nuclear polarization is tiny even at very low temperatures. To solve
this problem, the rst NMR measurements of T have been performed using
an optically pumped NMR (OPNMR) technique to increase the nuclear spin
polarization prior to the NMR detection [7, 103]. We found that for multiple-
QW samples standard NMR provides enough sensitivity for detection of
NMR signals at low temperatures T 10 K and in magnetic elds B
4 T. This chapter describes standard gallium NMR measurements on sample
M242 in the QHE regime
1
and focuses on ndings around Landau level
lling factor = 1.
As described in the preceding chapters, the key to understanding most of
the phenomenology of the 2DES near = 1 is tied to viewing its properties
dictated by the presence of Skyrmions - charged objects with an underly-
ing spin texture encompassing many reversed spins. What are the issues
pertaining to Skyrmions that NMR measurements endeavor to elucidate?
The most fervently pursued goal is the 2D electron spin polarization peak
at = 1, in T(T, ) experiments [7]. It is generally accepted that the 2D
electron spin polarization peak at = 1 unequivocally indicates the presence
of Skyrmions in the electronic ground state. In principle, measurements of
T(T, ) could easily determine the dependence of the size of Skyrmions on
Zeeman energy [Sec. (2.3.1)]. According to theory [95], the number of re-
versed spins within a Skyrmion varies with the ratio g = E
Z
/E
C
between
the Zeeman (E
Z
= [g

[
B
B) and the Coulomb energy [E
C
= e
2
/(l
B
)].
Above a critical g value ( g
c
) it is expected that Skyrmion-like excitations
evolve into single spin-ips [95]. Here [g

[ = 0.44 is the g-factor of elec-


trons in bulk GaAs,
B
is the Bohr magneton, 13
0
, l
B
=
_
/(eB

)
is the magnetic length, and B

is the perpendicular component of the ap-


plied magnetic eld. A very interesting aspect of heat capacity experiments
described in the precedent chapter is the disappearance of the nuclear spin
contribution to the measured heat capacity of sample M280 at g
c
0.04,
which is interpreted as evidence for a transition from Skyrmions to single
spin-ip excitations. At the risk of repetition, we recall that this g
c
is some-
what smaller than g
c
= 0.054 expected from theoretical calculations for
an ideally thin, disorder-free 2DES with no Landau level mixing [23, 95].
Quite puzzling, optical experiments involving nuclear spins [63] revealed a
signicantly smaller value for g
c
, close to 0.01.
1
We observed a similar behavior in sample M280. These data are not included here.
4.1. OVERVIEW 147
One of the objectives of NMR experiments described below is to clarify the
problem of the range of g over which Skyrmions are the stable excitations
of the = 1 QHE ground state. To this end, we studied the evolution of T
near = 1 in tilted magnetic elds up to 17 T, corresponding to an increase
of E
Z
and g by a factor of 3. The signature of Skyrmions, i.e., the 2D
electron spin polarization peak at = 1, is observed for g 0.022, whereas
for g 0.037 this feature disappears.
Well-established NMR probes of the electronic properties of bulk ma-
terials, such as the nuclear spin-lattice relaxation rate T
1
1
, may also pro-
vide crucial information about Skyrmions
2
. For example, a rapid nuclear
spin-lattice relaxation near = 1 would reect the presence of low-energy
electronic spin excitations associated to Skyrmion dynamics. We have used
the standard NMR technique to measure the low-T
71
Ga T
1
1
around = 1.
In doing so, besides assessing the fast nuclear-spin lattice relaxation near
= 1, new phenomena in the 2DES were exposed, as described below.
The present NMR experiments reveal important dierences around = 1
as compared to OPNMR results [7, 103]. The latter technique observes
a well developed (approaching full polarization T = 1) 2D electron spin
polarization peak at = 1 together with a strong suppression of T
1
1
due to
the single-particle spin (Zeeman) gap. In our NMR data, the 2D electron
spin polarization peak is signicantly smaller and no variation of T
1
1
around
= 1 is detected. The latter observation conrms the presence of rapidly
relaxing nuclei even at 1 detected in heat capacity experiments on
samples from the same wafer [9]. We discuss to what extent these results
can be explained in terms of electron density inhomogeneities across the
sample.
The organization of the present chapter is as follows. Section (4.2) is
devoted to essentials of NMR. First, some general elements of NMR are
presented. Then, we delineate the basic theoretical formalism useful in un-
derstanding the present experiments. Section (4.3) is devoted to the exper-
imental method. First, a brief account of the experimental setup is given
[Sec. (4.3.1)], followed by a discussion of how data were collected and ana-
lyzed [Sec. (4.3.2)]. Section (4.4) contains our results and interpretation of
the data. In Sec. (4.4.1) we specialize on our untilted magnetic eld data.
We discuss and compare these data to OPNMR results. In Sec. (4.4.2) we
consider electron spin polarization results in tilted magnetic elds approach-
ing the large Zeeman energy limit. Conclusions are given in Sec. (4.5).
2
Commonly, the geometrical arrangement of atoms in a crystal is termed the lattice.
In NMR, lattice refers to all degrees of freedom of the sample, except the nuclear ones.
148 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.1: (Top) (a) The NMR probe tuning/matching circuit. (b) Illus-
tration of the sample and the RF coil. (Bottom) Comparison of Gaussian
[Eq. (4.6)] and Lorentzian [Eq. (4.5)] line shapes for = 2.354 and
0
= 0.
4.2. PRINCIPLES OF NMR 149
4.2 Principles of NMR
4.2.1 General aspects of NMR
NMR refers to the case of a sample containing ^
n
nuclei of spin I and mag-
netic moment
n
I (where
n
is the nuclear gyromagnetic ratio). Usually, it
is assumed that the nuclear spin system is immersed in an applied static mag-
netic eld (which direction denes the z-direction in the laboratory frame,
B = Be
z
) and its energy level populations are described by the Boltzmann
distribution. The interaction energy between the applied static magnetic
eld and the nuclear magnetic moment is E
n
=
n
m
I
B =
0
m
I
,
where m
I
= I, I 1, . . . , I and
0
=
n
B is the nuclear Larmor pul-
sation. The thermal equilibrium condition for the nuclear magnetization
M (M
x
, M
y
, M
z
) is M
z
= M
0
and M
x
= M
y
= 0, where
M
0
=
^
n

2
n

2
I(I + 1)
3k
B
T
B. (4.1)
NMR measurements are basically the manipulation of the nuclear spins
in the externally imposed static magnetic eld B, by a time-varying radio-
frequency (RF) magnetic eld B
RF
. The heart of the NMR probe is a tuned
LC resonant circuit with the sample contained in the RF coil (a conventional
solenoid) [Fig. (4.1)]. The NMR coil produces the driving B
RF
by means of
which the energy separation between the nuclear energy levels is measured.
If B
RF
s frequency is close to
0
, transitions between nuclear energy levels
will occur and the nuclear spin system will absorb electromagnetic energy.
A simple, classical way of looking at the NMR is the following. The
nuclear magnetization vector precesses around the applied static magnetic
eld according to the well-known formula
dM
dt
=
n
MB. (4.2)
The angle between M and B is constant in time (t); its variation takes place
only with emission or absorption of energy. A superposed B
RF
, applied at a
90-degree angle to M, causes a forced precession of the nuclear magnetization
with decreasing latitude if its frequency is close to
0
. In a nuclear induction
experiment, B
RF
is turned o once the nuclear magnetization is pointing in
the xy-plane. Then, the free precession of the in-plane nuclear magnetization
induces a measurable voltage in the NMR coil.
150 CHAPTER 4. SKYRMIONS PROBED BY NMR
The phenomenological equations of Bloch
An ensemble of free nuclear spins is just an idealized test ground for NMR.
Next, we need to address the question of NMR detection in realistic, exper-
imental situations. Several physical quantities are important in describing
the phenomenon of magnetic resonance in solids, but two of them are inti-
mately related to the microscopic details of the nuclear spin system.
First, in the presence of an applied static magnetic eld, the estab-
lishment of thermal equilibrium between the nuclear spin system and the
substance containing the nuclei will require an exchange of energy between
nuclei and their surroundings (the so-called lattice). That is, the nuclear
spin system does not instantaneously reach thermal equilibrium with the lat-
tice when the external static magnetic eld is switched on. The rate of the
progress of the longitudinal nuclear magnetization M
z
towards its thermal
equilibrium value is T
1
1
, the nuclear spin-lattice relaxation rate.
Secondly, if there were no interactions among the nuclear spins, there
would be no decay of the in-plane nuclear magnetization in a nuclear in-
duction experiment. However, the spin precession is a function of the local
magnetic eld. For each single nucleus, because of nuclear dipole-dipole
interactions, the local magnetic eld is composed of the external static mag-
netic eld plus contributions from magnetic elds induced by the rest of
nuclear magnetic moments. The randomness from the dipole-dipole interac-
tions is the main source of the decay of the in-plane nuclear magnetization.
The nuclear spin-spin relaxation rate T
1
2
represents the rate of the decay
of transverse components of the nuclear magnetization (M
x
and M
y
).
In the phenomenological theory of Bloch [1], the components of the nu-
clear magnetization relax proportionally with their deviation from the ther-
mal equilibrium values
dM
z
dt
=
M
0
M
z
(t)
T
1
,
dM
x
dt
=
M
x
T
2
,
dM
y
dt
=
M
y
T
2
.
(4.3)
Combination of Eqs. (4.2) and (4.3) leads to Bloch equations in the labora-
4.2. PRINCIPLES OF NMR 151
tory system
dM
z
dt
=
n
(MB)
z
+
M
0
M
z
(t)
T
1
,
dM
x
dt
=
n
(MB)
x

M
x
T
2
,
dM
y
dt
=
n
(MB)
y

M
y
T
2
.
(4.4)
The macroscopic approach to NMR consists of solving the Bloch equa-
tions in the presence of a perturbing, linearly polarized RF eld B
RF
=
2B
1
cos t, applied along the x-axis. Lets consider that we are looking at
an absorption-type NMR experiment and we are interested in the power
absorbed in the RF coil, which is given by P
abs
= M
x
(dB
x
/dt). If the
RF excitation is a small perturbation, the response of the system may be
assumed proportional to it and can be written as M
x
= 2B
1
[
t
() cos t +

tt
() sin t], where
t
() and
tt
() are the real and imaginary parts of the
RF susceptibility () =
t
() i
tt
(). One can easily nd that the power
absorbed by the nuclear spin system is proportional to the imaginary part
of the complex susceptibility: P
abs
= 2B
2
1

tt
(). From the Bloch equations
we obtain
tt
() = (/2)
0

0
I(), where
0
= M
0
/B is the static nuclear
susceptibility and I() is the shape function (normalized to the unity) of
the NMR spectrum
I() =
T
1
2

1
(
0
)
2
+T
2
2
. (4.5)
The NMR line shape and T
1
2
In general, the function I() is a bell-shaped curve with a maximum at
0
.
As we can see in the calculations of the NMR spectrum using Bloch equa-
tions, T
1
2
is responsible for the broadening of the signal which would be a
-function if there were no nuclear spin-spin interactions. A noteworthy fea-
ture is that the exponential decay of the transverse magnetization results in
a Lorentzian shape of the NMR spectrum. While Lorentzian line shapes are
commonly found in experiments performed at high T, Gaussian line shapes
are characteristic for low temperatures. If T
1
2
= , the NMR spectrum of
Gaussian shape is described by
I() =
1

2
exp
_

(
0
)
2
2
2
_
. (4.6)
152 CHAPTER 4. SKYRMIONS PROBED BY NMR
Accordingly, the decay of the transverse magnetization M
+
= M
x
+iM
y
is
M
+
(t) = M
+
(0) exp (
2
t
2
/2). (4.7)
It is most convenient to characterize NMR line widths by specifying their
relation with T
1
2
. Our denition for the full width at the half maxi-
mum (FWHM) of the NMR signal (in the -domain) is = 2T
1
2
for
a Lorentzian line shape and = 2.36T
1
2
for a Gaussian line shape. At
the resonant frequency
0
, the quality factor Q is the ratio between
0
and
. The Lorentzian and Gaussian line shapes are shown for comparison in
Fig. (4.1)(bottom) for = 2.354 and
0
= 0. In order to characterize a
line shape to the full, one can study its variances (moments) dened as
()
n
=
_
(
0
)
n
I() d. (4.8)
For example, the second moment of a Gaussian curve is ()
2
=
2
.
Nuclear dipole-dipole couplings
A variety of seemingly eects, including the nuclear spin diusion and the
nuclear spin-spin relaxation, are results of the same basic interaction, namely
the dipole-dipole couplings of nuclear spins. We wish to elaborate now more
fully on these concepts as they are frequently used in this thesis.
A typical value of the nuclear spin-spin relaxation time in solids is
100 s, whereas the nuclear spin-lattice relaxation time can be equal
to seconds, minutes, hours or more. Consequently, the establishment of
thermal equilibrium between the nuclear spin system and the lattice can be
thought as proceeding in two steps
3
. First, the nuclear spin system reaches
an internal equilibrium at a temperature T
S
with a characteristic rate T
1
2
.
Secondly, T
S
tends towards the lattice temperature T with a characteristic
rate T
1
1
.
We take up now the subject of the nuclear spin diusion. In the presence
of nuclear dipole-dipole interactions it is possible to transport the nuclear
magnetization by ip-op transitions
4
. If we conne ourselves to an ac-
count of ip-op transitions of nearest neighbors, we obtain a nuclear spin
3
In the preceding chapter we have assigned a temperature T
S
to the nuclear spin system
without questioning if it is conceptually correct.
4
The dipolar Hamiltonian for A
n
interacting, like nuclei is
1
d
=

2

2
n
2
N
n
X
i=j=1

I
i
I
j
r
3
ij
3
(I
i
r
ij
)(I
j
r
ij
)
r
5
ij

, (4.9)
4.2. PRINCIPLES OF NMR 153
diusion constant D
n
Wa
2
, where W is the probability (per unit time) of
a ip-op transition of a pair of nearest identical nuclei and a is the distance
between them. A rough estimate of W for a cubic crystal is W T
1
2
/30.
The time required to transport the nuclear magnetization over a distance
r is of the order of r
2
/D
n
. Over distances long compared to the few lat-
tice spacings, the nuclear magnetization evolves according to a macroscopic
diusion equation
5
.
4.2.2 Interactions of nuclei with electrons. NMR in metals.
Since straight resonance experiments in conventional 2DESs are hardly prac-
ticable one may well ask what is the incentive for studying QHE systems by
NMR. The reason is that electron-nucleus interactions result in remarkable,
measurable eects such as the shift of the magnetic resonance, which, in
turn, yield valuable insights into the properties of 2D electrons. In the fol-
lowing we shall discuss the conceptual framework for NMR studies of QHE
systems, i.e., the interactions of nuclei with electrons.
The Knight shift (K
S
)
The general expression of the Hamiltonian for a nucleus at the position R
j
which interacts with its surroundings is [1]
H
j
(t) =
n
I B
j
(t), (4.10)
where B
j
(t) = B
j
) + B
j
(t) is the time-varying, total magnetic eld at
the position of the nucleus, B
j
) represents its average value, and B
j
(t)
stands for its uctuation. The rst term in the expression of B
j
(t) is re-
sponsible for the nuclear level splitting, whereas the second one induces a
time dependence in the expectation values of various nuclear properties.
In NMR measurements, transitions are induced at the resonance frequency
where r
ij
is the distance between nuclei i and j. To count the interaction between every
pair of nuclei only once there is a factor
1
2
and the double sum runs over the indices i and
j separately, excluding the value i = j. The homonuclear dipolar Hamiltonian contains
terms proportional to I
+
i
I

j
+ I

i
I
+
j
, which connect (m
I
, m
I
) to states (m
I
1, m
I
+ 1)
or (m
I
+ 1, m
I
1) by two simultaneous spin ips. These ip-op transitions conserve
the total Zeeman energy of the nuclear spin system and act as a means of transporting
the nuclear magnetization through the lattice.
5
Such a diusion equation was used in conjunction with magneto-thermal experi-
ments on sample M242 involving the cold nuclei of GaAs QWs and the hot nuclei of
Al
0.3
Ga
0.7
As barriers. There, we have tacitly supposed that the static applied eld in the
z-direction is homogeneous and T
1
1
is essentially zero.
154 CHAPTER 4. SKYRMIONS PROBED BY NMR

n
=
n
B
j
). Henceforth, as
n
is known, the measured resonance frequency
gives B
j
). The small dierence between B
j
) and B, due to the electronic
environment, denes the average value of the magnetic hyperne eld B
hf
j
)
and the magnetic hyperne shift tensor

K as
B
j
) = B+B
hf
j
) = B[1 +

K]. (4.11)
The magnetic hyperne shift has contributions from both the motion of
electrical charges and from the magnetic moments associated to the electron
spin. For the sake of simplicity, here we shall limit the discussion to the
Knight shift, which originates from the isotropic Fermi contact interaction,
i.e., the interaction between the nuclei and the conduction electrons inside
the nuclear volume. The Fermi contact interaction contribution to B
hf
j
) is
B
F
j
)
CGS
=
8
3

e
S
e
(R
j
)), (4.12)
where S
e
(R
j
) is the total electron spin operator at the position R
j
of the
nucleus. In the presence of an applied static magnetic eld along the z-axis,
the Knight shift K
S
is obtained as
B
j
)
z
= B +B
F
j
)
z
= B[1 +K
S
]. (4.13)
In a metal, the components of S
e
(R
j
) could be expressed in terms of one-
electron wave functions
k
(r) = U
k
(r)e
ikr
, where U
k
(r) = (V/)
1/2
u
k
(r)
is a function with the periodicity of the lattice normalized to the samples
volume, u
k
(r) is the periodic part of the Bloch state of wavevector k normal-
ized to the atomic volume, V is the volume of the sample, is the atomic
volume, and r is the electron coordinate. Only the z-component of S
e
(R
j
)
has a non-zero average value, which is given by
S
e
(R
j
))
z
=
1
2

k
[U
k
(R
j
)[
2
[n(E
k,
) n(E
k,
)], (4.14)
where n(E
k,
) represents the probability of nding a spin-up electron in a
state of energy E
k,
and it is given by the Fermi-Dirac distribution. As the
average value of the Fermi contact interaction is taken over all electronic
variables, the electrons in complete shells contribute nothing to it. Assum-
ing that the expectation values of the hyperne couplings are the same for
all the unlled orbitals near the Fermi surface and
e
is the electron spin
susceptibility, the (fractional) Knight shift writes as
K
S

B
B
=
8
3

e
[U
k
F
(R
j
)[
2
. (4.15)
4.2. PRINCIPLES OF NMR 155
The Knight shift is given as a relative value, namely the NMR line shift di-
vided by the reference line position (in % or ppm) [101]. When quoting the
value of the Knight shift, the reference with respect to which K
S
is computed
should be specied. The fractional Knight shift is nearly T-independent
in common metals and, provided that the hyperne eld coupling con-
stant /
hf
[usually dened by /
hf
= (8/3)[u
k
F
(R
j
)[
2

e
] is known, it
is ordinarily used to measure the electron spin susceptibility per site
e
as
K
S
= /
hf

e
/(
e
).
The nuclear spin-lattice relaxation rate (T
1
1
)
In metals, the spin-ip scattering of conduction electrons by the nucleus
constitutes an important channel for nuclear-spin lattice relaxation. In this
process both spins of the conduction electron and the nucleus are ipped.
The Fermi contact interaction Hamiltonian H
F
j
(t) =
n
I B
F
j
(t) contains
terms of the form I

S
e

(R
j
, t) which produce simultaneous electronic- and
nuclear-spin ips. Here S
e

= S
e
x
iS
e
y
are the (time-dependent) transverse
components of the total electron spin operator at the nuclear position R
j
.
A simple calculation, assuming non-interacting electrons, shows that T
1
1
in
a common 3D metal is
T
1
1
CGS
=
64
3
9

2
n

2
e

3
[U
k
F
(R
j
)[
4
[T(E
F
)]
2
k
B
T, (4.16)
where T(E
F
) is the DOS at the Fermi level for one direction of the electron
spin only. Korringas law tells us that T
1
1
is directly proportional to the
temperature, independent of the applied static magnetic eld, and simply
related to the fractional Knight shift
TK
2
S
=

4k
B
_

n
_
2
T
1
1
. (4.17)
Korringas law is usually taken as a representative of metallic behavior. It is
expected to be valid for metals whose dominant contribution to the magnetic
hyperne shift and T
1
1
comes from s-state electron-nucleus interactions. It
should be noted that electron-electron interactions aect
e
and T(E
F
).
Hence, the Korringa relation is modied to [113]
TK
2
S
=

4k
B
_

n
_
2
_
T
0
(E
F
)
T(E
F
)

e
0
_
2
T
1
1
, (4.18)
where T
0
(E
F
) and T(E
F
) are the DOS at the Fermi surface for an ideal
Fermi gas and for the real metal, respectively, and
e
0
and
e
are the corre-
sponding electron susceptibilities.
156 CHAPTER 4. SKYRMIONS PROBED BY NMR
A convenient expression for T
1
1
caused by the Fermi contact interaction,
which takes into account the interactions between electrons, follows from
the microscopic description of the nuclear spin-lattice relaxation processes.
Nuclear transitions are induced by the uctuations of the local hyperne
eld, which directly reect electronic spin uctuations [113]
T
1
1
=

2
n
2
_

[cos (
n
t)]B
F
+
(R
j
, t)B
F

(R
j
, 0)) dt, (4.19)
where ) means average over the states of the electronic system, repre-
sents the symmetric product, and B
F

(R
j
, t) S
e

(R
j
, t). The uctuation-
dissipation theorem relates the Fourier transform of the local hyperne eld
correlation function to the dissipative part of a generalized, frequency de-
pendent magnetic susceptibility
+
=
t
+
+i
tt
+
as [113]
_

[cos (
n
t)]B
F
+
(R
j
, t)B
F

(R
j
, 0)) dt

tt
+
(R
j
, R
j
,
n
)
1 e

n
/(k
B
T)
. (4.20)
Without sacricing any of the informational content of Eq. (4.19), one
can take the limit
n
k
B
T and use the spatial Fourier transform of

tt
+
(R
j
, R
j
,
n
) to obtain
T
1
1

k
B
T

n
_
1
(2)
2
_

tt
+
(q,
n
) dq
2
_
. (4.21)
Thus, the problem of deducing T
1
1
is reduced to that of calculating

tt
+
(q,
n
). Note that in the above formula we have explicitly considered a
2D Fourier transform to point out that the theory also applies to 2DESs [5].
The electric eld gradient (EFG)
No mention has been made so far of another important property of GaAs
QWs nuclei, namely the electrical quadrupole moment Q, which gives a
measure of the lack of sphericity in the distribution of the electric charge
inside the nucleus. The Coulomb interaction of the electrical quadrupole
moment with the surrounding electron charge distribution is described by
the Hamiltonian
H
Q
=
eQ
6I(2I 1)

,
V

_
3
2
(I

+I

I
2
_
, (4.22)
where and denote the directions in an arbitrary frame of reference
(X, Y, Z), I

and I
2
are operators associated to the angular momentum of
4.2. PRINCIPLES OF NMR 157
the nucleus, V

=

2
V (R
j
)

are the components of the local electric eld


gradient (EFG) tensor, and V (R
j
) is the electrostatic potential created by
the electrons at the position of the nucleus R
j
. The EFG tensor is symmetric
and by a change of the frame of reference can be diagonalized. In terms of
the principal axes (x, y, z), we have
H
Q
=
eQ
6I(2I 1)

(3I
2

I
2
). (4.23)
Furthermore, as

= 0, the EFG tensor is completely described by


two parameters. (As a consequence, for lattice sites of cubic symmetry, i.e.,
spherically symmetric charge distributions with V
xx
= V
yy
= V
zz
, the EFG
vanishes and there are no quadrupole interactions.) Generally, one chooses
to describe the EFG tensor by the largest principal component V
zz
and the
asymmetry parameter = (V
xx
V
yy
)/V
zz
. Within this convention, the
quadrupole Hamiltonian writes as
H
Q
=
h
Q
2
_
I
2
z

I
2
3
+

3
(I
2
x
I
2
y
)
_
, (4.24)
where
Q
is the quadrupole frequency

Q
=
3eQV
zz
2I(2I 1)h
. (4.25)
In the presence of an external static magnetic eld, which direction makes an
angle with the principal axis of the EFG tensor, the rst order quadrupole
coupling leads to the following nuclear energy levels
E
n,Q
=
n
m
I
B +
eQV
zz
4I(2I 1)
_
3 cos
2
1
2
_
[3m
2
I
I(I + 1)] (4.26)
where we have assumed that = 0. The quadrupole coupling shifts the
bare nuclear Zeeman energies by an amount which is related to both
Q
and . For I = 3/2 and = 0

the NMR spectrum is composed of a


central line at the frequency
0
/(2) and two satellites, equally shifted to
higher [
0
/(2) +
Q
/2] and lower [
0
/(2)
Q
/2] frequency. The central
resonance is the m = 1/2 1/2 transition, whereas the satellites are the
m = 1/2 3/2 transitions.
4.2.3 NMR in QHE systems
K
S
in GaAs quantum wells
We have seen in the preceding section the basic theory behind NMR experi-
ments and how NMR can be used to probe the electron spin polarization of
158 CHAPTER 4. SKYRMIONS PROBED BY NMR
a metallic system. In this section we point out some usage and denitions
specic to NMR studies in QHE systems. We reiterate that NMR uses the
nuclear magnetic moment
n
I to measure the additional local magnetic
eld B
hf
(R
j
, t) produced by the electrons at the position R
j
of the nucleus.
The hyperne coupling Hamiltonian [1] is
H
hf
j
(t) =
n
I B
hf
(R
j
, t). (4.27)
The average value of the hyperne eld B
hf
j
) is measured as the shift of the
NMR line [compared to the reference line position in the absence of B
hf
j
]
K
S
(R
j
) [s
1
] =

n
2
B
hf
j
)
z
, (4.28)
where it is assumed that the applied static magnetic eld lies along the z-
direction. We remark here that the shift denition given in Eqs. (4.28) and
(4.31) diers from the usual one [ Eqs. (4.11) and (4.15)]. That is, for 2DES
and QHE, the NMR shift is a measure of the local electron spin polariza-
tion and not of the electron spin susceptibility per site; the absolute value
of the shift [given by Eq. (4.28) in frequency units] measures directly the
time-averaged value of B
hf
(R
j
, t). The hyperne eld includes contribu-
tions from several sources, the strengths of which can be quite dierent. For
GaAs, calculations show that at the point of the rst conduction band, the
s-like contribution dominates the total electron charge-density distribution,
which is localized on the nuclei [88]. Thus, the isotropic Fermi contact inter-
action gives the dominant contribution to the hyperne eld. The hyperne
coupling Hamiltonian [Eq. (4.27)] reduces to
H
F
j
(t) I S
e
(R
j
, t)
e
(R
j
), (4.29)
where
e
(R
j
) is the local value of the electron density and S
e
(R
j
, t) is the
electron spin density at the nuclear site. The corresponding NMR line shift,
called the Knight shift as in simple metals, directly measures the local spin
polarization T(R
j
)
K
S
(R
j
) T(R
j
)
e
(R
j
). (4.30)
For multiple-QW samples, the nuclei in the barriers, where the electron
density is vanishingly small, will provide the reference line position. Strictly
speaking, some non-zero electron density regions also exist in barriers close
to QWs and Si-doping layers. However, these regions cover only small per-
centage of the volume of the barriers, so they contribute to the negligible
part of the NMR signal from barriers and are not expected to modify its
4.2. PRINCIPLES OF NMR 159
peak position. This is indeed conrmed experimentally, as the barriers peak
position depends neither on the temperature nor on the recovery time t
R
[see Sec. (4.3.2) below], and it is found at the same position as the signal
from the substrate.
To infer the global polarization T(T, ) from the measured Knight shift,
we shall assume a spatially homogeneous electron spin polarization. The
validity of this assumption resides in the fast dynamics of electrons along
the QWs over the NMR observation time [93]. Thus, under the simplest
assumption of uniform electron density (
e
= n/w), the nuclei in GaAs
QWs provide a single NMR line whose shift measures directly the electron
polarization
K
S
(T, ) = A
c
n
w
T(T, ). (4.31)
The above relation denes the eective hyperne coupling constant A
c
. In
fact, as the electron density is not constant across the QW, its z-dependence
somewhat aects the NMR line shape
6
. However, as much as the
e
(z) is
a rigid function, essentially independent of T and , the line shape will also
be rigid and we can use the shift of the peak position in Eq. (4.31) to deduce
T(T, ) [93]. The dierence between the actual
e
(z) and the average value
n/w, will be taken into account by A
c
.
The value of A
c
dened in Eq. (4.31) is experimentally determined from
the NMR shift obtained at low T, high B, and at lling factors (such as
= 1/3) where the 2DES is fully polarized (T = 1). Using the value
for
71
Ga determined in this way in OPNMR experiments [55, 64], A
c

=
4.5 10
13
cm
3
/s, we calculate from Eq. (4.31) a maximum shift in our
sample of K
1=1
S
= 24 kHz and deduce the spin polarization of 2D electrons
as T(T, ) = K
S
(T, )/K
1=1
S
. We have an uncertainty of 10% in our
samples w, resulting in a similar uncertainty in our calculated K
1=1
S
. The
experimental uncertainty of the measured K
S
(T, ) is about 10 20%.
T
1
1
in GaAs quantum wells
The Fermi contact interaction is responsible for the nuclear spin-lattice re-
laxation in common metals, such as copper [1, 94]. This relaxation mecha-
6
In previous OPNMR electron spin polarization studies [55, 64]
e
(z) was approximated
by a sinusoidal shape and an intrinsic hyperne shift for the nuclei in the center of the QW
was deduced. Corrections to 1 [Eq. (4.31)] arising from realistic approximations to
e
(z)
are negligible [93]. We also note that our denition of 1 [Eq. (4.31)] makes no distinction
between the nuclei close to the GaAs/Al
0.3
Ga
0.7
As interface and those located near the
centers of the QWs. Such a distinction may be important when NMR spectra are sensitive
to dynamical processes.
160 CHAPTER 4. SKYRMIONS PROBED BY NMR
nism involves a simultaneous spin ip of both spins of the conduction elec-
tron and the nucleus, and energy conservation requires that there must be
closely spaced spin and electron levels with energy comparable to the
nuclear Zeeman gap. For a 2DES, however, in the presence of a quantiz-
ing magnetic eld and in the absence of disorder, the single-particle energy
spectrum is discrete. The ip-op process under discussion is forbidden and
no Korringa relaxation is expected. Adding to the puzzle is the fact that
the nuclear spin-lattice relaxation in pure insulating crystals is negligible
small through direct interaction between the nuclear spins and the lattice
degrees of freedom. The nuclear spin-lattice relaxation actually observed in
high-purity bulk GaAs has been interpreted as magnetic relaxation, owing
to nuclear spin diusion to paramagnetic centers, which serve as intermedi-
ates between nuclear spins and lattice
7
. The nuclear spin diusion has no
B-dependence and may add a constant background to T
1
1
(T, B), dictated
by the relaxing impurities. In the presence of high magnetic elds, how-
ever, paramagnetic impurities, if there are any, should be frozen into their
lowest spin state; thus, the nuclear spin-lattice relaxation by paramagnetic
impurities is expected to be suppressed.
Central to any discussion of the nuclear spin-lattice relaxation in
electron-doped GaAs QWs under the QHE regime is the existence of nearly
gapless electron spin-ip excitations
8
. Our understanding of a nite
T
1
1
is in terms of additional degrees of freedom present in real 2DESs.
These degrees of freedom could be generally ascribed to impurities (disor-
der) [5, 52, 105] and for the particular case of 1 or 1, also to the
presence of Skyrmions in the electronic ground state [24, 44]. Calculations
of T
1
1
in the QHE systems, which properly take into account the eects of
disorder and electron-electron interactions, appeared well before the begin-
ning of the Skyrmion era [5, 105]. By invoking the presence of Skyrmions,
recent theories [24, 44, 87] explain the observed one thousand fold enhance-
ment of T
1
1
at [ 1[ 0.1, compared to the zero magnetic eld value.
The theoretical predictions for T
1
1
(T) near = 1 have been summarized in
Chapter 2.
7
An interesting nuclear spin-lattice relaxation process occurs in conventional ferromag-
nets such as cobalt. Here, the nuclear spin-lattice relaxation is by nuclear spin diusion
away from the domain walls and ultimately to the lattice by coupling with the conduction
electrons. Consequently, the T-dependence of the observed T
1
1
does not follow Korringas
law.
8
We note that at B = 0 our 2DES is metallic-like. A simple calculation of T
1
1
, based
on the analogy with the 3D free electron gas and on the assumption that the Fermi contact
interaction is the only relaxation mechanism yields T
1
1
= 9
3
A
2
c
(n/w)
2
(k
B
T/E
2
F
). For
sample M242 at T = 1 K, we obtain T
1
1
4 10
4
s
1
.
4.3. EXPERIMENTAL METHODS 161
We wish now to complete the discussion of T
1
1
by examining the inte-
ger QHE regime, where the discreteness of the 2D single-electron spectrum
manifests in an interesting way in the B- and T-dependence of the nuclear
spin-lattice relaxation rate [105]. Because the energy conservation in the ip-
op process can be fullled at any nite T by the participation of a phonon,
an activated T-dependence of T
1
1
is expected at low temperatures. How-
ever, at high T, because the number of thermally excited phonons increases
linearly as the temperature is increased, the phonon-assisted T
1
1
should fol-
low the usual Korringas law. On the other hand, since the disorder leads to
a broadening of Landau levels (with DOS displaying maxima at half-integer
lling factors), one expects that T
1
1
undergoes magneto-quantum oscilla-
tions, similar to that of the longitudinal magnetoresistance. This oscillatory
behavior of T
1
1
has been observed via electron spin resonance (ESR) of 2D
electrons
9
by Berg et al. [11]. The experimentally determined T
1
1
never
become smaller than T
1
1
= 1 10
3
s
1
, whereas the value of T
1
1
at max-
ima (half-integer lling factors) is about 4 10
3
s
1
. This value is one
order of magnitude larger than that expected at zero magnetic eld.
4.3 Experimental methods
4.3.1 Experimental setup and the sample
In order to increase the NMR signal, two pieces of the multiple-QW sample,
each 25 mm
2
in area, were placed together (tightly wound with Teon
tape) into the RF coil [Fig. (4.1)(top)]. Thus, the total number of QWs
nuclei (Ga or As) contained in our measured specimen was 310
18
, which
was sucient for the observation of
69,71
Ga NMR signals at temperatures
below 10 K and magnetic elds above 4 T with our NMR apparatus [79].
The experimental setup consists of a single-coil duplexing scheme, i.e., the
same coil is used for transmitting the RF pulses and receiving the nuclear
induction signals. The RF coils were conventional solenoids, constructed
by hand from copper wire wound around a metallic slab whose cross-section
9
Coupling as they do the 2DES and the nuclear spin system of GaAs QWs, the nuclear-
electron interactions can manifest themselves in the study of either system. The electronic
spin splitting is not solely due to the Zeeman eect, but also inuenced by the nuclear-spin
polarization, which acts via hyperne interaction on the electronic spins. In thermody-
namic equilibrium, the nuclear polarization is negligible even at very low T. However, the
nuclear spins may be dynamically polarized. In the presence of a large nuclear polariza-
tion, the ESR line will be shifted. The T
1
1
of nuclei could be observed via ESR [28] of
2D electrons because a non-equilibrium nuclear polarization results in a non-equilibrium
ESR line position, which relaxes towards the equilibrium with a characteristic rate T
1
1
.
162 CHAPTER 4. SKYRMIONS PROBED BY NMR
was slightly larger than the samples. By potting the coils in epoxy resin, the
sample could easily be inserted and replaced. This conguration also min-
imizes the magneto-acoustic ringing [38]. We list in Table (4.1) important
parameters of RF coils used in the present study.
Table 4.1: Parameters of two coils used in NMR experiments are given. All
sample coils have an internal cross-section of 1 5 mm
2
. Typically, RF
coils have a quality factor of about 100 at room temperature, measured
near the bottom end of the resonance frequency [f
0
=
0
/(2)] range. The
corresponding B-range is given for
71
Ga. Note that with the use of both Ga
isotopes we were able to cover an extended range of B with the same coil.
Coil f
0
-range (MHz) B-range (T) Number of turns Length (mm)
# 1 49 - 101 3.8 - 7.8 24 7.5
# 2 108 - 228 8.3 - 17.5 10 6.5
The NMR spectra were recorded on a custom built pulsed NMR spectrom-
eter. The temperature control was provided by a standard variable temper-
ature insert (VTI) down to 1.5 K (in pumped liquid
4
He or in a gas ow
above 4.2 K), and by a
3
He/
4
He dilution refrigerator down to 0.05 K. The
NMR system is equipped with a superconducting magnet with maximum
eld strength of 17 T (in pumped liquid
4
He). The magnetic eld spatial
homogeneity over the dimensions of the sample is better than 10 ppm and
its temporal stability is few ppm per hour. An important advantage of our
standard VTI setup was the possibility of varying in situ the tilt angle
between B and the normal to the plane of the 2D electron layers, together
with the low-T tuning and matching of the RF coil (the so-called bottom
tuning)
10
. This enabled us to vary at will three parameters: T, , and
B (above 1.5 K and up to 17 T). For very low-T measurements, the RF
coil with xed tuning/matching was mounted into the mixing chamber of
the dilution refrigerator. This ensured good thermal contact to the sample
and kept optimal NMR sensitivity, but restricted each low-T experiment to
single T-dependence at xed and B.
10
The resonant frequency of the circuit containing the coil is tuned by adding in series
a variable capacitor C
tune
. The impedance can be adjusted to 50 (real) by adding in
parallel a capacitor C
match
[Fig. (4.1)(top)] or inductor [38].
4.3. EXPERIMENTAL METHODS 163
4.3.2 Experimental technique
Essential to pulsed NMR experiments, is the detection of the free induction
decay (FID). By applying a strong RF pulse, an alternating RF magnetic
eld is produced along the axis of the RF coil. This RF magnetic eld, whose
frequency satises the condition for nuclear resonance, turns the nuclear
magnetization out of the z-direction. When the nuclear magnetization has
turned by 90-degrees, one speaks of a /2-pulse. The RF magnetic eld is
switched o at the end of the time interval corresponding to the /2-pulse
and the NMR coil detects the transient voltage due to the precessing nuclear
magnetization around z-axis. That is, the FID experiment probes the free
precession of the transverse nuclear magnetization following a /2-pulse.
It is however clear that the transverse nuclear magnetization will decay
with a rate T
1
2
due to variations in the local magnetic eld experienced
by the nuclei. The origin of these variations may be inhomogeneities of the
applied static magnetic eld or internal inhomogeneities, such as the dipolar
magnetic elds of neighboring nuclei. If T
1
2
is very large, it is dicult to
see the NMR signal because the FID will disappear before the instrumental
dead time ( 2 s) is over.
Spin echoes are commonly used to reform the NMR signal some time
after the /2-pulse in order to avoid the instrumental dead times which
always corrupt the start of the FID. The spin echo sequence (/2
echo) [15, 46] is pictorially illustrated in Fig. (4.2). The result of a /2-pulse
is a non-equilibrium conguration of nuclear magnetic moments. After the
/2-pulse, the nuclear spins precess freely with dierent frequencies imposed
upon them by the existing inhomogeneities. As they dephase with time, the
macroscopic magnetization of the nuclear spin system is lost. However,
application of a -pulse after a time causes all nuclear spins to come back
together at the same rate that they precessed prior to the -pulse. After a
time 2 all nuclear spins are again in phase and the echo is produced.
NMR measurements
The rst obvious problem with the NMR in multiple-QW samples is to sep-
arate the signal of QWs from the barriers signal, which is about 7 times
larger, and from the even stronger signal of the substrate
11
. In our experi-
ments, since the nuclear polarization builds up because of nuclear spin-lattice
11
Note that in OPNMR this is done automatically, as one pumps only the QWs
nuclei [1, 7]. We also stress out that standard NMR experiments dier from OPNMR
measurements in that they do not require the realization of a highly polarized nuclear
spin system through nuclear dynamic polarization.
164 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.2: (Top) Pulse sequence used to detect the NMR spectra. (Bottom) Schematic diagram of the spin-echo
sequence [/2)
x
)
y
echo] in the rotating frame. The spin echo is essentially two FIDs back-to-back.
4.3. EXPERIMENTAL METHODS 165
relaxation processes, the unwanted signals can be suppressed by exploiting
the dierence between the T
1
1
of QWs nuclei and T
1
1
of nuclei located in
the GaAs substrate and barriers. At low temperature, T
1
1
of the substrate
is extremely small due to the absence of mobile electrons and dopants.
Figure 4.3:
71
Ga NMR spectra taken at (a) f
0
= 73.915 MHz with t
R
=
256 s (top) and 2 s (bottom), and at (b) f
0
= 192.052 MHz with t
R
= 128
s (top) and 32 s (bottom). For short recovery times (lower spectra) the
contribution is essentially from nuclei in QWs, while for longer times (upper
spectra) barriers signal becomes stronger than the one from QWs.
For barriers nuclei, T
1
1
is increased because of the -doping; the largest
T
1
1
is normally expected for QWs nuclei due to 2D electrons (except for
the gapped phases of the 2DES at very low T). Thus, at the beginning of
our pulse sequence, the nuclear magnetization of the Ga isotope of interest is
set to zero by a comb of /2-pulses. After the magnetization has recovered
during time t
R
, we perform standard spin-echo or FID read out and obtain
166 CHAPTER 4. SKYRMIONS PROBED BY NMR
spectra by Fourier transforming [Figs. (4.3), (4.4), and (4.5)]. At short t
R
(1 10 s), the signal contains mainly the QWs contribution; for longer
recovery times ( 10 s), it is dominated by the barriers.
Figure 4.4:
71
Ga NMR spectra measured at = 1, = 0

, and indicated
temperatures. For short recovery times (lower spectra) the contribution is
mainly from nuclei in QWs, while for longer times (upper spectra) barriers
signal becomes dominant. The NMR frequency shift is given relative to the
position of the barriers line.
4.3. EXPERIMENTAL METHODS 167
The quantity called K
S
of Ga nuclei in QWs is here dened as the dier-
ence between the peak positions of QWs and barriers. As the same NMR
shift data were obtained both by FID and by the spin-echo sequence, only
the latter technique is used in the present work.
The complete pulse sequence used for the present measurements is de-
picted in Fig. (4.2)(top). The /2-pulse lasts for times of the order of 1 s
which provide uniform irradiation over a wide spectral range. Typical sep-
aration between the /2- and -pulses was

= 100 s. The pulse spacing
within the comb was set to 3 ms. The signal-to-noise is improved by aver-
aging over several NMR scans and a phase cycling is executed during the
accumulation of spectra to eliminate spurious eects.
A brief comment on the possibility of heating the sample is in order
here. First, the 2DES is extremely susceptible to heating by eddy currents
around = 1
12
. Secondly, a direct coupling between the RF pulses and
the 2DESs might be possible [64]. In our experiments, any variation of the
temperature of the 2DES during the measurement is easily detected through
the T-dependence of the measured NMR shift. The amount of RF heating is
determined by the RF-pulse power and the repetition rate of the spin-echo
sequence. Values of these parameters are chosen to avoid any possible RF
heating. That is, our standard test to denitely exclude the problems of
heating consists of checking that results do not change when the RF-pulse
power is reduced by an order of magnitude.
The quadrupole splitting of the resonance
Our measurements were carried out using both Ga isotopes. As expected
from Eq. (4.28), we indeed nd that
71
K
S
/
69
K
S
=
71

n
/
69

n
, and for sim-
plicity all the results are normalized to
71
Ga isotope using the
71

n
/
69

n
=
1.27 ratio.
12
We note that if the electrons are heated by the Joule eect, no spin depolarization
is expected at = 1 as the electron heating is minimal (the longitudinal resistance is
essentially zero). The ohmic heating, however, eectively depolarizes the spins of the
electronic systems for small variations of the lling factor away from = 1 by increasing
the 2DES temperature over the lattice temperature. This phenomenon, which is known in
magneto-transport experiments as the breakdown of the QHE, has a single-particle nature
and it is independent of the existence of Skyrmions [62].
168 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.5: Set of NMR spectra comparing the line shapes of
71
Ga and
69
Ga,
measured at T = 1.5 K, = 3/4, and = 0

. The lower (upper) spectra


are taken with short (long) recovery time t
R
, respectively. The quadrupole
splitting is better observed at short t
R
, as shoulders to the central QW line.
The NMR frequency shift is given relative to the barriers line position.
Typical
69
Ga and
71
Ga NMR spectra are illustrated in Fig. (4.5). For
the QW line we can clearly observe the quadrupole splitting [Figs. (4.3)(a)
and (4.5) (bottom traces)]. On both sides of the main line appear satellites
whose splitting for two isotopes scales with known electrical quadrupole mo-
ments,
69
Q/
71
Q = 1.6 [Table (A.1)]. Such quadrupole eects can occur only
if the local symmetry is not cubic. One may suspect that breaking of the
cubic symmetry is induced by the slight mismatch of the lattice constants
of GaAs and Al
0.3
Ga
0.7
As, where the barriers impose their lattice constant
to the in-plane lattice constant of QWs. The quadrupole coupling is clearly
resolved at high temperatures [78], indicating that the eect is homogeneous
over all QWs in the sample. The strain-splitting ratio in GaAs QWs was
measured to be (4.7 0.4) 10
6
kHz
1
for
75
As [45]. From our exper-
4.3. EXPERIMENTAL METHODS 169
imental conditions [Fig. (4.3)(a), bottom trace] we deduce an estimate of
9 10
5
for the actual strain in our sample, close to residual value 6 10
5
found in Ref. [45]. At low T, the satellite splittings are obscured by the line-
broadening, but they do contribute to the total width of the QWs resonance
line. Note that no quadrupole splitting appears in the barriers line.
71
Ga T
1
1
measurements
The same RF pulse sequence described above is used for T
1
1
measurements
and is called the saturation recovery method, i.e., we measure the NMR
signal intensity as a function of t
R
. A typical result, presented in Fig. (4.6),
singles out the diculty to dene the saturation value of the NMR signal at
long t
R
. The point is that at very long recovery times the barriers signal
prevails in the NMR spectrum and thus pollutes QWs signal. In order to
minimize this eect we analyzed only the (integral over the) left half of the
QWs resonance line. Our analysis of the nuclear spin-lattice relaxation rate
is based on Eq. (4.3) for all investigated T- and -ranges. Fits to the Bloch
relaxation formula [Eq. (4.3)] reproduce well the measured recovery of the
nuclear magnetization, suggesting a single dominant source of relaxation.
That is, within our experimental accuracy determined by the signal to noise
ratio and the possibility to separate the signal of QWs from the partially
overlapping barriers, we did not nd evidence for a multi-exponential re-
laxation. We also mention that for t
R
10 s, the nuclear spin diusion could
be considered as an ecient process for the thermal relaxation, ensuring a
homogeneous spin temperature across the QWs. Hence, a single T
1
1
which
does not vary along z-axis is a satisfactory approximation in describing the
growth of QWs nuclear magnetization in our NMR experiments.
Reliable T
1
1
measurements become impossible at higher temperatures;
the NMR shift becomes small and the overlap with the barriers line is too
strong, leading to ecient nuclear spin diusion between QWs and barriers
which modies the NMR line shape
13
. Finally, we wish to point out that
for the gapped phases of the 2DES at very low T, T
1
1
can be extremely
small and the time for nuclear magnetization to reach its equilibrium value
becomes prohibitively long for T
1
1
measurements.
13
We observed that the barriers signal grows proportional to T
1
1
in the QWs. The
question should be raised then as how the contribution of the barriers nuclei to the
measured heat capacity near = 1 is usually not observed. We believe that the barriers
nuclear contribution remains hidden in heat capacity experiments because the thermal
relaxation is essentially completed when the nuclear spin diusion into the barriers sets
in.
170 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.6: Recovery time t
R
dependence of the NMR signal measured at
T = 1.5 K, = 1 and = 0

. The dashed curve is an one-parameter


exponential t to the data [Eq. (4.3)] used to determine T
1
1
.
The NMR line shape
At this point of the discussion we would like to briey remark on the NMR
line shape. The QWs line shape is observed at short recovery times; it is
rather symmetric and, neglecting the weak contribution from the barriers,
can be approximated to a Gaussian. At intermediate times, the NMR line
shape has signicant contributions from both QWs and barriers. Conse-
quently, the separation of the overlapping QWs and barriers peaks is more
involved and less reliable. The barriers signal dominates the NMR spec-
tra in the limit of very long recovery times and has the expected Gaussian
shape. The FWHM values of the QWs and barriers Gaussian resonances
are summarized in Table (4.2). The FWHM value of the NMR signal of
barriers that emerges from our analysis is f
B
= 3.75 0.28 kHz, in good
agreement with 3.5 kHz obtained in OPNMR experiments. The broaden-
4.3. EXPERIMENTAL METHODS 171
ing of the barriers signal is attributed to the dipolar coupling between the
nuclear spins. We determine for the NMR signal of QWs a FWHM value
f
QW
15 kHz, four times bigger than f
B
. This value is also three times
larger than f
QW
observed in OPNMR experiments. The observed broad-
ening is likely to originate from the quadrupole interaction of the nuclei with
the EFG as f
QW
is approximately equal to our experimental quadrupole
splitting.
Table 4.2: The measured FWHM (in kHz) of the NMR signal of barriers
(f
B
) and QWs (f
QW
) at T = 1.5 K and = 0

is given for selected


B-values (in units of T).
B 4.7 5.7 6.7 7.6 9 11 13 15
f
B
3.8 3.5 3.9 3.8 3.5 3.4 3.9 4.4
f
QW
12.9 14.9 15.3 15.4 14.3 15.5 15.3 16.7
As discussed in Appendix B, one can obtain useful information about the 2D
electron spin dynamics from the measurements of the QWs line width as a
function of T [55, 64]. In the present study, our attention was not directed
to the changes in the QWs line width with the temperature. An obvious
diculty we have confronted is that the analysis of the QWs line shape in
our sample can only be made if the quadrupole eects are properly taken
into account, which is laborious or rather impossible.
Finally, we note that insight into the dependence of NMR line shape on
the nuclear polarization prole along the z-axis [
n
(z)],
e
(z), and electron
spin polarization and dynamics around = 1 has been attained recently by
Sinova et al. [93]. These authors have found a tremendous sensitivity of the
spectral line shape to
n
(z) which explains the inhomogeneous broadening
(asymmetry) of the QWs line shape observed in OPNMR experiments. In
general,
n
(z) not only depends on the experimental technique but also
on microscopic mechanisms (such as the nuclear spin diusion and nuclear
spin-spin relaxation). The question should be raised then as to how the
scenario described in Ref. [93] accounts for the nearly symmetrical QWs
line shape observed in our experiments [bottom traces in Fig. (4.4)]. It
could simply mean that the standard NMR favors a atter
n
(z) compared
to the OPNMR method.
172 CHAPTER 4. SKYRMIONS PROBED BY NMR
71
Ga T
1
2
measurements
Using a conventional (

2
) spin-echo method, T
1
2
could be easily deter-
mined: the echo intensity is monitored in a sequence of (

2
) spin-echo
measurements, each (

2
) spin-echo measurement taken with a dierent
value of the total time delay (t), separating the /2 pulse from the echo.
Our denition of T
1
2
is based on the assumption of a Gaussian decay of the
echo intensity
I
echo
(t) exp
_

1
2
_
t
T
2
_
2
_
. (4.32)
Figure (4.7)(a) displays a typical time dependence of the echo intensity
which directly yields T
1
2
. Under our experimental conditions,
71
T
1
2
does
not seem to depend neither on B nor on T (including dilution refrigerator
temperatures). We determined
71
T
1
2
= 1.38 0.15 ms
1
[Fig. (4.7)(b)].
It is noticeable that
71
T
1
2
was previously estimated from the FWHM of
the barriers signal in OPNMR measurements: a Gaussian FWHM of 3.5
kHz gives
71
T
1
2
= 9.32 ms
1
. This calculated
71
T
1
2
is usually expected to
be larger than the one obtained from spin-echo experiments. Interestingly
enough, spin-echo standard NMR measurements reveal a
71
T
1
2
about 7
times smaller than the value estimated from the FWHM of the barriers
signal in OPNMR measurements.
The present
71
T
1
2
value allows an estimate of the nuclear spin diusion
coecient for
71
Ga nuclei of
71
D
n
0.7410
13
cm
2
/s. Given the fact that
71
D
n
/
75
D
n
= 0.7, we infer for
75
As nuclei
71
D
n
1.0510
13
cm
2
/s, in ex-
cellent agreement with previously reported
75
D
n
110
13
cm
2
/s [84]. We
could condently assert that in our sample it takes about 10 s to transport
the nuclear magnetization over a distance r = 100

A. Finally, we mention
that such T
1
2
measurements may prove fruitful in studying the motion of
Skyrmions
14
.
14
The key feature of T
1
2
experiments is that each (

2
) spin-echo measurement probes
the irreversible dephasing of the transverse nuclear magnetization during the time interval
t. A good example of physical situation where one can take advantage of this feature of
the experimental procedure is the QHE ferromagnet near = 1, where Skyrmions are
accommodated in the electronic ground state. Consider rst the situation where Skyrmion-
size uctuations are frozen over the (varying) time window t/2, separating the /2
pulse from the pulse. In this case, a static contribution (due to the Skyrmions) to the
local magnetic eld is experienced by the nuclei. This time-independent inhomogeneity
is eliminated in a (

2
) spin-echo measurement, i.e., the -pulse refocuses all nuclei
which dephased during the time interval t/2. On the contrary, if the uctuations in the
size of Skyrmions are fast over the time interval t/2, the result is a time-dependent local
magnetic eld inhomogeneity. In this case, only a fraction of the nuclei which dephased
4.3. EXPERIMENTAL METHODS 173
Figure 4.7: (a) Typical time dependence of the echo intensity in a
71
T
1
2
experiment. Referring to Eq. (4.32), we use a semilogarithmic plot of I
echo
vs t
2
to extract
71
T
1
2
(inset). (b)
71
T
1
2
vs B at T = 1.5 K and = 0

.
174 CHAPTER 4. SKYRMIONS PROBED BY NMR
4.4 Results and discussion
4.4.1 Untilted magnetic eld data
The electron spin polarization
I shall describe our results in the light of previous OPNMR experiments [7,
55, 103] and start with the electron spin polarization data collected at = 0

as a function of and T. I must stress that our method has the great
advantage that T can be measured at xed and T over a broad -range.
Therefore, Fig. (4.8) is the rst fairly complete picture of the electron spin
polarization in the fractional QHE regime at = 0

and T = 1.5 K, from


NMR measurements. Also displayed in Fig. (4.8) are selected low-T data
points at = 1, 1/2, 1/3, where full polarization is expected for T 0.
We observe a weak maximum at = 1 followed by a monotonous increase
of T as decreases, without any pronounced structure for 1/3. The
increase of T is roughly linear in B, however, the behavior changes at very
low temperatures, as illustrated by the T = 0.1 K data points. The T-
dependence of T taken at these lling factors [Fig. (4.9)] conrms that for
both = 1/2 and 1/3 the full polarization is reached for T 0.
On the other hand, our measurements revealed an unexpected T-
dependence of T at = 1 [Figs. (4.9)(b) and (4.10)(a)]. The electron spin
polarization is only slightly increased as we cool down the sample from 1.5 K
to 0.1 K [T( = 1) 0.4], meaning that the corresponding electron spin po-
larization peak remains very weak. Figure (4.11) shows the signicant eect
of the temperature on the electron spin polarization peak
15
. At T = 4.2 K
the electron spin polarization peak is completely lost
16
. In OPNMR exper-
during the time interval t/2 will be refocused by the application of the pulse and the
echo intensity would be drastically reduced. By measuring T
1
2
, lets say, as a function
of T, it is therefore possible to address the microscopic motion of Skyrmions. It is worth
pointing out that frozen Skyrmions might be too strongly perturbed by /2 pulses.
To observe the dynamics of Skyrmions then one should devise modied pulse sequences
and/or use small tip-angle RF pulses.
15
For T 1.5 K, note that the amplitude of the electron spin polarization peak is small
and comparable to the measurement inaccuracy. In principle, the electron-hole symmetry
in the lowest Landau level is reected in a symmetric shape of the electron spin polarization
peak [7, 103]. Given the experimental uncertainty, this issue cannot be addressed by the
present experiments.
16
The very pronounced T-dependence of the electron spin polarization peak is similar to
the behavior observed in magneto-absorption experiments on highest-quality single-QW
heterojunctions [see Ref. [73] and Fig. 3 (sample B) in M.J. Manfra, B.B. Goldberg, L.N.
Pfeier, and K.W. West, Optical Determination of the Spin Polarization of a Quantum
Hall Ferromagnet, Physica E 1, 28 (1997)].
4.4. RESULTS AND DISCUSSION 175
iments on comparable samples (in our case g = 0.014 at = 1 and = 0

),
the electron spin polarization peak is clearly visible already at T 4 K [7, 56]
and at T = 1.5 K it reaches nearly full polarization with its at 4% -wide
top [Fig. (4.10)(b)]. We note that at 1.2 and T = 1.5 K both NMR and
OPNMR techniques nd T 0.20 0.25 [Fig. (4.10)(b)], i.e., considerably
less than unity. Finally, an interesting feature of the T(T, ) data shown
in Fig. (4.11) is its strong T-dependence observed at = 3/4 compared to
very small variations measured at = 0.9.
Figure 4.8:
71
Ga K
S
(left axis) and T (right axis) vs B at = 0

. The
corresponding is given on the top axis. Knight shift data were collected
at T = 1.5 K on
71
Ga () and on
69
Ga (), which is renormalized to
71
Ga
K
S
by the ratio
71
/
69
= 1.27. Data points at T = 0.1 K are denoted by
lled triangles (). The QHE state at = 1 corresponds to B = 5.7 T.
176 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.9: T (right axis) and
71
Ga K
S
(left axis) vs T. (a) Open diamonds
denote T(T) at B 14.8 T and = 40

( = 1/2). The solid curve is a


single-parameter t to the data (see text). (b) Open circles denote T(T) at
B 17 T and = 0

( = 1/3). The dashed curve is a two-parameter t


to the data. Open squares denote T(T) at B 5.7 T and = 0

( = 1).
4.4. RESULTS AND DISCUSSION 177
Figure 4.10: (a) K
S
vs T at = 1 in NMR () and OPNMR () experi-
ments [7]. OPNMR data is taken at = 28.5

and B = 7.05 T. (b) NMR


() vs OPNMR () K
S
() results at T = 1.5 K. OPNMR experiments ob-
tain K
S
() by tilting the sample in a xed magnetic eld B = 7.05 T. All
NMR data is taken at = 0

. Curves are calculations explained in the text.


178 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.11:
71
Ga K
S
(left axis) and T (right axis) vs B at = 0

and
T = 1.5 K (), 2.2 K (), and 4.2 K (). The corresponding is given on
the top axis. A closed triangle () is used for the 0.1 K data point at = 1.
T(T) at = 1/2 and = 1/3
We pause here to briey discuss our measured T(T) at = 1/2 and
= 1/3 [Fig. (4.9)]. As a rst remark we mention that, due to the ex-
perimental uncertainty
17
, in the vicinity of = 1/3, the low-T limit of T
17
Note that the data points at low T [Fig. (4.9)(b)] have a larger and asymmetric error
bar, showing possible greater consistency with K
P=1
S
= 24 kHz. In our experiments, there
are two factors which introduce errors near = 1/3 and underestimate K
S
at low T. While
the broadening of the QWs resonance line is responsible for the larger uncertainty, the
asymmetry in the error bar is due to the excessively long t
R
necessary for the observation
4.4. RESULTS AND DISCUSSION 179
seems to be 10% lower than the saturation value observed at = 1/2
[Fig. (4.9)]. The T-dependence of T at = 1/3 is shown by open circles
in Fig. (4.9)(b). As expected, we observe that T(T 0; = 1/3) 1.
Although there is a signicant scatter at low-T [78], T(T; = 1/3) agrees
well with OPNMR results [55, 64, 78]. The raw Knight shift data is tted to
K
sat
S
( = 1/3) tanh(
1/3
/4k
B
T), yielding K
sat
S
( = 1/3) = 20 2 kHz and

1/3
= 1.6E
Z
, in agreement with the OPNMR result
1/3
= 1.82E
Z
[55, 64]
and the theoretical estimate
1/3
2E
Z
[71]. Note that, according to the-
ory, the contribution of single-particle excitations to T dominates at low-T
near = 1/3 [71], in contrast to = 1 where the collective (spin-wave)
modes are expected to control T at nite T [54].
In Fig. (4.9)(a) we present the measured T-dependence of the electron
spin polarization at = 1/2 in our sample. We note that T reaches the full
polarization value as T 0, implying that the ground state of the 2DES is
fully spin polarized at lling factor one half. Near half-integer lling factors
( = 1/2, 1/4, . . . ), a large body of experimental and theoretical results can
be cast in a surprisingly simple picture of non-interacting composite fermions
(CFs) [29, 51] with the same charge and spin as electrons
18
. An important
issue in the physics of CFs is the spin polarization of the 2DES at half-
integer lling factors. In an eort to understand the spin polarization of the
2DES close to = 1/2, Park and Jain [85] introduced a phenomenological
parameter, the CF polarization mass m

p
, which is proportional to the ratio
of the cyclotron and Coulomb energies. These authors obtained an estimate
for m

p
at = 1/2
m

p
/m
e

= 0.60
_
B

[T], (4.33)
where m
e
is the electrons mass in vacuum. The parameter m

p
, combined
with a parabolic dispersion law for CFs at = 1/2, uniquely determines the
electron spin polarization at any given T and B. Quite remarkably, our T-
dependence of T at = 1/2 supports the non-interacting CF picture. Lets
assume that CFs have a g-factor roughly the same as electrons and consider
parabolic bands occupied by n CFs with mass m

p
. Hence, the density of
states D

(E) (for spin-up and spin-down CFs) is D

(E) = D

(EE
Z
/2),
where D

= m

p
/(2
2
), and is the step function. Making use of Fermi-
of the saturated peak position of the barriers.
18
In the CF model, the Coulomb interaction of one electron with all others is replaced
with a Chern-Simons gauge eld, equivalent to attaching an even number (2m) of ux
quanta to each electron. In the mean eld approximation the gauge eld equals the
external magnetic eld at = 1/(2m) and the system of interacting electrons is replaced
by one of non-interacting CFs in zero average magnetic eld.
180 CHAPTER 4. SKYRMIONS PROBED BY NMR
Dirac distribution we nd
T(T, B) =
2D

k
B
T
n
_
_
E
Z
2k
B
T
tanh
1
_
1 +
exp(
n
D

k
B
T
)
sinh
2
(
E
Z
2k
B
T
)
_

1
2
_
_
. (4.34)
Depending on the strength of the magnetic eld, this model predicts either
partially or completely polarized Fermi sea of CFs in the T 0 limit,
i.e., T(T = 0) = minD

E
Z
/n, 1. Taking m

p
as a tting parameter, in
Fig. (4.9)(a) we show the best t of Eq. (4.34) to the data by the solid
curve. This curve indeed provide a reasonable description of the data at
low temperatures (T 2 K), which are of experimental relevance for the
physics of CFs. At higher T, the measured T falls below the calculated
T(T, B). The deduced m

p
= 1.7m
e
value is found to be in good agreement
with the polarization mass predicted by Eq. (4.33): m

p
/m
e
= 2.0 at B =
B

= 11.4 T.
The nuclear spin-lattice relaxation rate
Figure (4.12)(a) presents the -dependence of T
1
1
measured at T = 1.5 K.
At this temperature essentially no variation is found over the investigated
-range, with the exception of = 1/2 and = 1/3 points where T
1
1
is
signicantly smaller. In particular, the value at = 1 is the same as values
nearby, i.e., we observe no decrease of the nuclear relaxation rate. This is
in contrast with OPNMR results [Fig. (4.12)(b) and Appendix B] reporting
strong reduction of T
1
1
(for 1.5 K T 4.2 K) in the vicinity of = 1,
which is indeed expected as a signature of a gapped QHE state at = 1.
On the other hand, the NMR value of T
1
1
at = 1 is in good agreement
with our heat capacity experiments performed on a sample from the same
wafer [9]. These measurements revealed that up to 70% of QWs nuclei have
large T
1
1
as they contribute to the measured heat capacity within the ex-
perimental time scale. Furthermore, the T
1
1
() found in these calorimetric
measurements [Fig. (3.21)] corroborates with our NMR data. This behavior
could be accounted for by sample inhomogeneities and disorder which in-
troduce localized states in the gap and compressible islands in the 2DES,
leading to additional relaxation channels for the QWs nuclei
19
. The area
occupied by the compressible regions can be estimated by comparing the
19
Since T
1
1
is expected to be very dierent between compressible and incompressible
parts of the sample, standard NMR observes rst the nuclei with large T
1
1
, located near
the compressible regions.
4.4. RESULTS AND DISCUSSION 181
Figure 4.12: (a)
71
Ga T
1
1
vs B at T = 1.5 K and = 0

. The corresponding
is given on the top axis. The uncertainty in our measured T
1
1
is 20%. (b)
NMR () vs OPNMR () T
1
1
() results near = 1. OPNMR experiments
obtained T
1
1
() by tilting the sample in a xed magnetic eld B = 7.05 T.
182 CHAPTER 4. SKYRMIONS PROBED BY NMR
jump in the electron orbital magnetization in the integer and the fractional
QHE regime [112]. This yields an upper bound of only 10% of the total
area of the sample, however, by the mechanism of nuclear spin-diusion,
the percentage of nuclei aected by rapid relaxation may be much larger.
In any case, we have searched on purpose at long recovery times t
R
for the
NMR signal arising from slow-relaxing nuclei in incompressible regions near
the predicted T = 1 position. Nothing was detected; if some slow relaxing
NMR signal existed, it must involve a very small amount of the observed
nuclei. In our mind, this apparent discrepancy could indicate the presence
of a nite, -dependent density of magnetic states in the gap which provide
ecient relaxation channels for the QWs nuclei. It is at present impossible
to know the origin and the nature of these states: they could be extrinsic
as magnetic impurities, or intrinsic as formation of local moments or more
exotic states in the disorder potential.
Regarding the comparison of OPNMR [103] and NMR T
1
1
data, we
also remark that both techniques are in reasonable agreement away from
= 1 [Fig. (4.12)(b)]. Recall that within the disorder-free description of
the 2DES, the increase of T
1
1
away from the = 1 is attributed to the
Skyrmion-mediated nuclear-spin lattice relaxation. For example, at T =
1.5 K and 0.85 one has T
1
1
(OPNMR) 0.05 0.01 s
1
[103] and
T
1
1
(NMR) = 0.0760.015 s
1
. Moreover, at 0.9 both techniques reveal
an approximately T-independent T
1
1
between T = 1.5 K and T = 4.2 K
(not shown here).
Discussion
We begin the examination of the = 0

data by recalling that, according


to previous theoretical and experimental work [7, 14], the electron spin po-
larization peak at = 1 [Fig. (4.10)(b)] signals the presence of Skyrmions
in the QHE ground state. Theoretical modelling of the electron spin polar-
ization around = 1 [solid curve in Fig. (4.10)(b)], by Brey et al. [14], has
found convincing agreement with OPNMR results. Our NMR data agree
with this numerical simulation for llings away from = 1 (i.e., 1.2
and 0.8), but disagree at 1. The etiology of this discrepancy is
discussed in detail below.
Going a step further, one of the crux ndings of OPNMR experiments
is the T-dependence of T at = 1 [Fig. (4.10)(a)], which received outstand-
ing theoretical justication. Recent work by Song et al. [96], performed by
a similar OPNMR technique, has uncovered a similar T-dependence of T
at = 1. As small variations in have a dramatic eect on T, the most
4.4. RESULTS AND DISCUSSION 183
prosaic reason for small discrepancies between theory and (OPNMR) exper-
imental data is the diculty of precisely locating = 1 in the investigated
samples [53]. Even admitting this exceptional circumstance, none of the ex-
isting theoretical results can be trivially reconciled with our measured T(T)
at = 1 in the T 0 limit [open squares in Figs. (4.9)(b) and 4.10(a)].
In other words, while full polarization of the 2DES is always predicted at
low-T, our measured T(T) at = 1 falls well below the limit imposed by
T(T) for localized, non-interacting electrons [dashed line in Fig. (4.10)(a)].
Clearly, a subtle physical mechanism is at work here. It is a signicant
point that, at high-T (T 6 K), our T(T) data corroborates with previous
OPNMR T(T) measurements (and theory).
Lets turn now our attention to T
1
1
measurements. Our data [closed
circles in Fig. (4.12)] manifestly disagree with the OPNMR results around
= 1 [Fig. (4.12)(b)]. Moreover, since T
1
1
0 as T 0 at = 1/3
but does not at = 1, our nuclear spin-lattice relaxation rate data seem
to contradict magneto-transport observations, which nd a large QHE ex-
citation gap at = 1 compared to that at = 1/3. Nevertheless, at least
at a qualitative level, one might understand this apparent inconsistency. It
is quite simple to imagine how samples inhomogeneities tend to smear the
minimum in T
1
1
() at = 1. Physically, it is plausible to consider that the
T
1
1
at nominal lling factor = 1 in an inhomogeneous 2DES could be con-
structed from averaging over the T
1
1
contributions at = 1. Recounting
that T
1
1
increases linearly with [1 [ due to the presence of Skyrmions, it
is easy to see that the -averaged T
1
1
will have essentially no dependence
on the lling factor around = 1 and a fast nuclear spin-lattice relaxation
is observed down to very low T. At = 1/3 the variations in the electron
density will leave T
1
1
essentially unchanged because the elementary exci-
tations are not Skyrmion-like and there is no reason for an enhancement of
T
1
1
near = 1/3. Consequently, a slowing down of the nuclear spin-lattice
relaxation is observed as T 0 at nominally lling factor = 1/3.
Lets come back at T(T, ) measurements. It may appear at the rst
sight that samples inhomogeneities are a good candidate for explaining our
puzzling observations at = 1. It has been suggested recently [82] that the
presence of disorder modies the shape of the polarization peak at = 1.
These results [82] have been obtained in the particular case of weak, micro-
scopic disorder but naturally lead to a more general idea of inhomogeneities,
short-range and/or long-range, governing both the magnitude and shape of
the spin polarization peak at = 1. Motivated by the NMR and the heat
capacity T
1
1
results, we already mentioned the possibility of long-range
electron density inhomogeneities. Indeed, to minimize the contamination
184 CHAPTER 4. SKYRMIONS PROBED BY NMR
during the MBE growth the wafer was not rotated, and our measured speci-
mens present some electron density gradients. A conservative estimate gives
the maximum density variation of 5% over the entire sample surface. We
recall here that the peak in T() observed in OPNMR is a quite robust
characteristic, i.e., it is rather wide around = 1, covering more than 5% at
1.5 K. Thus, even if we admit that our NMR sample contains a span of 10%
in the electron density, signicant percentage of QW nuclei should observe
high electron spin polarization and should have a small T
1
1
. We should
add that the same conclusion is reached for any kind of static microscopic
electron density inhomogeneity of the same order of magnitude
20
.
One might object that nuclei in the high-T domains are particularly
dicult to observe by NMR using saturation-recovery method, because of
their very slow relaxation rate. This is true only at low-T; at T 4.2 K,
both in high-T and in low-T domains, nuclei are relaxing fast enough to
be easily observable. Therefore, at high-T, both NMR and OPNMR are
expected to give similar results on similar samples. However, in contrast to
the OPNMR results, no electron spin polarization peak is observed by NMR
at T = 4.2 K [Fig. (4.11)], and T
1
1
(NMR) is one order of magnitude larger
than T
1
1
(OPNMR) [Fig. (4.10)(b)].
A possible mechanism capable of diminishing the T( = 1) polarization
peak and hiding slow relaxation by some sort of average over might be a
dynamical (and necessarily short-range) inhomogeneity. That is, the system
is composed of spin-textured domains, whose -distribution moves rapidly
on the time scale of NMR data acquisition (i.e., faster than 10 s). Why
and how should such hypothetical situation be realized in our NMR case
and be absent in OPNMR experiments is not clear to us.
The discrepancy between the OPNMR and NMR results at = 1 may
also be induced by the dierences between the two experimental techniques.
One cannot rule out the possibility that the electron spin polarization peak
observed in OPNMR experiments might be aected by the optical pump-
ing process. Hyperne interaction with oriented nuclei changes the eective
electron Zeeman splitting; this Overhauser shift can signicantly alter the
20
The presence of signicant static microscopic disorder can be readily seen in magneto-
transport measurements: for example, the QHE excitations gaps would be markedly re-
duced. The measured QHE excitation gaps on samples from the same wafer are compara-
ble to those measured in high-quality conventional single-layer 2DESs, clearly discounting
the presence of strong, random disorder. One fundamental dierence between NMR and
magneto-transport measurements should be mentioned here. Transport QHE excitation
gaps are known to be sensitive to the percolation structure of the incompressible region
through the sample. While in transport experiments the compressible part of the sample
is quiet, in NMR (which is a thermodynamic measurement) it does matter.
4.4. RESULTS AND DISCUSSION 185
magnitude of the eective electron Zeeman energy (see Fig. 6 in Ref. [63])
and it decays with a very long time constant after the optical pumping pro-
cess is stopped [107]. Therefore, in OPNMR measurements, 2D electrons
might experience an articial g, leading to a larger electron spin polariza-
tion. Finally, we note that a study of the T( = 1) peak as a function of
the g factor has been performed using polarized magneto-reectivity
21
. In
these experiments, for QWs where [g[ < 0.4, T( = 1) saturates at low T
at values which are well below 100%, suggesting a strong spin mixing of the
two spin sublevels of the zeroth LL at small values of E
Z
.
4.4.2 Tilted-magnetic eld electron spin polarization
As aforementioned, the determination of the critical value g
c
of the Zeeman
to Coulomb energy ratio g = E
Z
/E
C
above which Skyrmions are converted
to single spin ips is an important test for the Skyrmion picture near = 1.
We have used the tilted-magnetic eld technique to tune g and examine the
eect of an increasing Zeeman energy on the = 1 electron spin polarization
peak. Figure (4.13)(a) includes results for T at three dierent total magnetic
eld values, measured by varying . In this experiment, for each B, the rst
data point is measured at = 0

while the largest investigated tilt angle


corresponds to = 2. To emphasize the behavior of T around = 1, the
data are displayed in Fig. (4.13)(b) as a function of . In this gure we also
included the data taken at = 0

by varying B [from Fig. 4.8)], as well


as the numerically calculated T() curve for the non-interacting electron
model [96] [dashed curve in Fig. (4.13)(b)]. The main sources of uncertainty
in are geometrical errors which arise from the angle measurement 1

.
We rst focus on the two nearly identical T() curves measured at B =
17 T and 14.8 T. The electron spin polarization drops to zero when 2,
as expected in the independent electron model, and closely mimics the non-
interacting electron picture for 1.2 2 [dashed line in Fig. (4.13)(b)].
The key feature of these two curves is the absence of T( 1) peak which
is replaced by a monotonic -dependence. In contrast to these data, the
polarization peak at = 1 is clearly seen in the B = 9 T data (where
g = 0.022), as well as in already presented = 0

data (where g = 0.014).


Regarding the low -range of our T = 1.5 K data, we mention that
21
V. Zhitomirsky, R. Chughtai, R.J. Nicholas, and M. Henini, Spin Polarization of 2D
Electrons in the Quantum Hall Ferromagnet: Evidence for a Partially Polarized State
around Filling Factor One, in Workbook of the 14
th
International Conference on the
Electronic Properties of Two-Dimensional Systems, (Prague, July 30 - August 3, 2001),
p. 17.
186 CHAPTER 4. SKYRMIONS PROBED BY NMR
Figure 4.13: (a) Tilt angle dependence of
71
Ga K
S
(left axis) and T (right
axis) measured at T = 1.5 K and B = 9 T (3), 14.8 T (), and 17 T ().
(b) Data points shown in panel (a) are displayed here as a function of .
Open and lled triangles denote = 0

data points [Fig. (4.8)]. The dashed


curve is T() at T = 0 for the independent-electron model (see text).
4.4. RESULTS AND DISCUSSION 187
because of dierent T-dependencies at various lling factors 1/3 2/3,
T() curves exhibit an apparent minimum between = 2/3 and = 1/3.
In fact, from our studies [78] we know that the 2DES at = 1/2 is fully
polarized in the low-T limit [see in Fig. (4.13)(b) the T = 0.1 K data point
measured at = 0

].
Before discussing the tilted-magnetic eld results, we make two impor-
tant remarks about the data presented in Fig. (4.13). First, as in the present
study tilt angles up to 80

were used, we warn the reader that strong tilt


of the sample with respect to the magnetic eld could induce signicant
modications of the electronic wave function, which in turn may inuence
the lowest-energy excitations of the 2DES [see for details Sec. (1.6.4)]. This
eect is not included in our analysis. Secondly, a crossing between the rst
spin-down Landau level and the second spin-up Landau level is predicted
22
to occur in our sample for 74

[see for details Sec. (1.6.4)]. For B = 17 T,


the highest investigated B, 74

corresponds to 1.2. Thus, this as-


pect does not concern the discussion of spin properties of the = 1 QHE
state. Furthermore, in our experimental T() data, up to B = 17 T, we did
not nd signatures of a Landau level crossing. It is the overall agreement
between the experimental data measured at B = 17 T and the theoretically
predicted T() for a non-interacting 2DES, which rst deserves attention.
The 2D electron spin polarization at = 2 is essentially zero at all B-values
up to the highest investigated tilt angle, and at B = 17 T T() is approach-
ing this point along the prediction for the free electrons. At = 1 and
T = 1.5 K, in the presence of total magnetic eld B = 17 T, we actu-
ally nd the 2DES polarized only to 70%. However, it is clear that high
magnetic elds and low temperatures push the observed T() dependence
towards the independent electron picture in the full -range [Fig. 4.13(b)].
In particular, this is corroborated by T = 0.1 K data points at = 1/3 and
1/2, given by lled symbols.
We next discuss the eect of large Zeeman energy on the 2D electron spin
polarization around = 1. For g 0.022 [open diamonds in Fig. 4.13(b)],
T() clearly displays a maximum at = 1, which we associated to the pres-
ence of Skyrmions in the electronic ground state. This is also suggested by
our very low-T tilted-eld heat capacity studies on a sample from the same
wafer. In these experiments, the nuclear contribution of Ga and As atoms
to the measured heat capacity near = 1, up to the highest investigated tilt
angle 53

( g 0.023), indicates that the nuclear spin-lattice relaxation


occurs via Skyrmions. Disappearance of the T( 1) peak in NMR data
22
T. Jungwirth and A.H. MacDonald (private communication).
188 CHAPTER 4. SKYRMIONS PROBED BY NMR
taken in magnetic elds of 14.8 T or higher is an experimental evidence
for a Skyrmion to single spin-ip transition near g
c
0.037. The data are
somewhat aected by the nite temperature at which the experiment was
performed and the range of investigated, xed B is not ne enough to convey
a precise value for g
c
. Nevertheless, given the fact that, in our experiments,
T near = 1 slowly approaches the low-T saturation value when T is low-
ered below 1.5 K, we believe that g
c
0.037 is a reasonable upper bound
estimate for the existence of Skyrmions near 1. Note that numerical
Hartree-Fock calculations for the present sample, taking into account only
the nite-thickness of the 2D electron layers, predict g
c
= 0.046 in reasonable
agreement with the experiment [23]. Based on calculations which incorpo-
rate both nite-width and Landau-level mixing corrections [60], we expect a
further reduction of g
c
, improving the agreement between the experimental
data and theory.
4.5 Conclusion
A primary objective of this work was to obtain information about the
= 1 QHE state from a standard NMR measurement in a multiple-QW,
GaAs/Al
0.3
Ga
0.7
As sample. Present 2D electron spin polarization data sup-
port the Skyrmion picture around = 1 at low magnetic eld, and provide
an upper bound estimate of the critical Zeeman energy (0.037 in units of
Coulomb energy) for the existence of Skyrmions. This value is in good
agreement with both theory and previous heat capacity measurements. In
the high magnetic eld and low temperature limit, the observed T() ap-
proaches the predicted T() for non-interacting 2D electrons over the full
range of Landau level lling factors 1/3 2. Our data at = 0

also
reveal that at = 1/2 and = 1/3 the 2DES is fully polarized as T 0.
We also discuss the observed dierence between the optically pumped
NMR data [7] and our NMR measurements. At = 1, we observe by
standard NMR a smearing out of the Skyrmion polarization peak, which
is complete at T = 4.2 K, and no reduction of the nuclear spin-lattice re-
laxation rate at low temperatures. The possibility that this is due to the
inhomogeneity of the electron density is analyzed, and we conclude that ob-
served features could be accounted for only if the electron density pattern
is not static. Even if this hypothesis were true, it remains unclear why it
is absent from low-T OPNMR data. The discrepancy between our results
and those obtained by optically pumped NMR might also be due to the
experimental technique.
Epilogue
i am plans mult, indca nimeni n-a fost gasit vrednic sa de-
schida cartea, nici sa se uite n ea.
Apocalipsa Sf. Ioan Teologul
It is a truism that principal results of modern physics reach far beyond
the domain in which they are won. The theoretical idea of Skyrmions was
rst advanced in particle physics. The new views with regard to quantum
Hall eect (QHE) ferromagnets and pseudomagnets
23
, which have risen
as a result of vigorous theoretical and experimental research on Skyrmions,
revolutioned our deeply held ideas about ferromagnetism an old condensed
matter physics paradigm. Yet nobody dares, perhaps, to prophesy that there
is no conceivable experience which can refute the reality of QHE Skyrmions.
Praising the fruits of this thesis is not a cheerful task. For me, the tri-
umphs of this work, if any, were the triumphs of the faith. Gallant eorts
sometimes have only a puny outcome; my joy was ruined by countless neg-
ative results of experimental enterprise. In the preceding chapters, in the
main, I have conned my account to our positive results of research, and
have not dealt in detail with problems which are not yet completely cleared
up. The impression might thus be given that the measurements presented
in this thesis left a denitive mark on the physics of QHE Skyrmions. That
is by no means the case.
23
The pseudospin degree of freedom is represented by discrete quantum numbers such
as the electronic spin or the Landau level index.
189
190 EPILOGUE
While several issues concerning QHE Skyrmions are now settled, there
still remain dicult problems, in particular, those related to the eect of
disorder. The basic theoretical understanding of the role of disorder for the
integer QHE is rather well-established. Nevertheless, describing the complex
behavior in disordered two-dimensional (2D) systems of strongly correlated
electrons has been a beckoning, yet elusive goal for the 2D physics com-
munity. Both microscopic and macroscopic inhomogeneities, always present
in macroscale (centimeter) samples, are potentially serious sources of error.
Inescapable, high ideals such as Skyrmions are besmirched by the existence
of inhomogeneities or swamped by the bulk eects. One might seek to
resolve this in measurements of thermodynamic properties on conventional
(single quantum well) samples. I believe that clever experiments on multiple-
quantum-well (QW) samples, however, will continue to guide our ceaseless
understanding of the QHE.
Additional research is needed to rene many of the topics addressed in
this thesis. In the following, I tentatively formulate a prospectus of future
experimental work. I shall start with ac heat capacity experiments. In at
least one important aspect, the present ac calorimetric study remains in-
complete. The usefulness of our method is revealed by the fact that the
heat capacity of two-dimensional electron systems (2DESs) could be mea-
sured over a large range of magnetic elds and tilt angles. Therefore, I claim
that tilted-magnetic eld ac heat capacity measurements, through a detailed
study of the shape of heat capacity oscillations in the integer QHE regime,
would provide relevant information on the functional form of both Landau
level broadening and spin-splitting. Our heat capacity experiments in the
integer QHE regime indicated a nite density of states between Landau lev-
els. This aspect, inherent to the presence of disorder, should be explored
experimentally more thoroughly.
Of all dicult moments of the Skyrmion crusading era, quasi-adiabatic
heat capacity measurements were the most impressive. These measurements
had cost thousands liters of liquid helium, they had tried the patience of
several people, and they had taught me that one could shine only scrimp
light on Skyrmion crystals. I hope that future heat capacity experiments
will elucidate the enigma of the Skyrmion liquid-to-solid phase transition.
Based on our present knowledge, high-density multiple-QW heterostructures
are believed the most promising candidates for the observation of the heat
capacity peak at low temperatures. It should be noted that several im-
pediments have left nuclear spin-lattice relaxation rate (T
1
1
) calorimetric
measurements with limited capacity to contribute new insights into the Skyr-
mion physics. A consistent framework for this type of experiments currently
EPILOGUE 191
exists, and future developments in this direction are desirable.
Another possible continuation of the present heat capacity work is the
study of two spatially separated 2DESs. The separation between layers can
be made small enough so that the inter-layer and intra-layer Coulomb in-
teractions are comparable in strength, i.e., strong correlations between the
layers are bring about. Bilayer 2DESs display complex physical phenomena,
which thus far have been revealed and studied only by means of standard
magneto-transport methods. It seems to be no reason to doubt that heat
capacity measurements on bilayer 2DESs will be among the most exciting
future research perspectives. On the experimental side, eort in this direc-
tion was already directed, by the growth of multiple-double-QW samples.
On such a sample, we have discovered evidence that Skyrmions are present
in the electronic ground state at total Landau level lling factor
t
= 2
( = 1 in each layer), consistent with a fully polarized QHE state in which
the spins in each layer are aligned parallel to the magnetic eld. An in-
teresting question is what happens around
t
= 2 as the layer separation
is reduced or the sample is tilted in the magnetic eld. On the theoretical
side, it was shown that at
t
= 2 a canted antiferromagnetic phase could
be observed under realistically obtainable experimental conditions
24
. Ad-
ditionally, in strong magnetic elds, double-layer 2DESs can form at
t
= 1
an unusual broken-symmetry state with spontaneous interlayer phase co-
herence. Whenever the system has spontaneous interlayer phase coherence
and is incompressible, its low-lying excitations are expected to be interesting
objects
25
, called Merons. The Meron is essentially one-half of a Skyrmion.
The present standard nuclear magnetic resonance (NMR) experiments
furnish a new test of the 2D electron spin polarization in the extreme quan-
tum limit. We have focused our eorts on comparing the spin polarization
data with existing theoretical models and with a series of experiments. It
is necessary to insist here that NMR results at = 1/2 support the non-
interacting composite fermion picture. The most intriguing results of our
observations remain the spin polarization and T
1
1
at = 1. The subtle in-
terplay between electron-electron interactions and disorder in the fractional
QHE regime prompted us to consider in detail the implications of samples
inhomogeneities on thermodynamic properties of the QHE ferromagnet at
24
S. Das Sarma, S. Sachdev, and L. Zheng, Canted Antiferromagnetic and Spin-Singlet
Quantum Hall States in Double-Layer Systems, Phys. Rev. B 58, 4672 (1998).
25
Moon K., H. Mori, K. Yang, S.M. Girvin, A. H. MacDonald, L. Zheng, D. Yoshioka,
and S.-C. Zhang, Spontaneous Interlayer Coherence in Double-Layer Quantum Hall Eect
Systems: Charged Vortices and Kosterlitz-Thouless Phase Transitions, Phys. Rev. B 51,
5138 (1995).
192 EPILOGUE
= 1, and ask which of various sources of disorder is responsible for the
observed behavior. We argued that our NMR data cannot be reconciled
with theory (and partially with optically pumped NMR and heat capacity
measurements) without express consideration of short-range, dynamical in-
homogeneities across the sample. Experimental work is now underway at the
Grenoble High Magnetic Field Laboratory; NMR measurements in the frac-
tional QHE were extended towards very low-T and new, interesting aspects
occurred
26
. Since I have said not much about NMR in the integer QHE
regime, I should mention that measurements on sample M280 suggest the
existence of interesting spin phenomena in the high Landau levels ( = 2 and
= 4). We probably need to fortify our understanding of the role of nuclear
spins in the integer QHE regime with direct, NMR T
1
1
measurements.
In Chapter 1, we have followed a common trend in the presentation
of transport coecients, and displayed the longitudinal resistance results
on an arbitrary scale. Seldom is the denition of the actual resistivity
used because this entails determining a geometric value for scaling the re-
sistance measurement. Such geometric factors are particularly dicult to
ascertain for our investigated multiple-QW specimens. The subtle and crit-
ical nature of the magnetotransport in 2DESs is beautifully reected in
hysteretic and anisotropic behaviors, spikes and nonlinearities in the longi-
tudinal resistance, and metastable fractional QHE states. These eects are
usually observed in measurements where the current ow is precisely con-
trolled through lithographic denition of the sample geometry. It should be
noted that some of these anomalies are expected to be present in our sam-
ples. Systematic magneto-transport studies on lithographically patterned
Hall bridges will certainly help to brighten the existing tableau of spin or
pseudospin spatial ordering at various lling factors such as = 4/3 or
= 2.
This thesis outlines a strategy to improve the links between dierent
levels of experimental description. Our heat capacity, NMR, and magneto-
transport experiments uncovered only a very small part of the world of
QHEs. These phenomena are rich enough in changing hues and patterns to
allure us to explore them in all directions.
26
N. Freytag, Y. Tokunaga, M. Horvatic, C. Berthier, M. Shayegan, and L.-P. Levy,
Observation of a New Phase Transition between Fully and Partially Polarized Quantum
Hall States with Charge and Spin gaps at = 2/3, preprint, cond-mat/0105590.
Appendix A
The NSS in GaAs QWs and
the Schottky Heat Capacity
Consider the case of a sample containing nuclei of spin I and nuclear gy-
romagnetic ratio
n
. Lets suppose that the sample is placed in a static
magnetic eld B. Then, the nuclear energy levels are equidistantly spaced
by
n
=
n
B. Here
n
/(2) is given in units of MHz/T and B in
units of T. Alternatively, we can express
n
in units of K as
n
=
[K]
n
/(2)[T
1
]B [T], where = (h/k
B
) 1 MHz = 4.8 10
2
mK.
The structure of nuclear levels is reected in the heat capacity in a
very interesting way [41]. At
n
k
B
T all levels will be well populated,
whereas at
n
k
B
T the upper levels will be scarcely occupied. For
n

k
B
T transitions between levels take place in appreciable amounts. This
rapid change in the internal energy corresponds to large specic heat which
becomes zero at both high and low temperatures. There is an intriguing
possibility of a hump (Schottky eect) in the measured C when
n
k
B
T,
which in general will be superimposed on lattice and other contributions.
Lets focus now on the nuclear spin system (NSS) of GaAs quantum
wells (QWs) and quantitatively explore the Schottky eect. Electromagnetic
properties of GaAs QWs nuclei are listed in Table (A.1). It is convenient
to invoke the canonical partition function for a I =
3
2
single nucleus
Z =
m=3/2

m=3/2
exp
_
m
n
k
B
T
_
, (A.1)
where m = 3/2, 1/2, 1/2, 3/2. Provided that m
n
/(k
B
T) is a small
number, it is permissible to make a linear expansion of the Boltzmann ex-
193
194 Appendix A: Schottky Heat Capacity
Table A.1: Published nuclear data for
71
Ga,
69
Ga, and
75
As: spin I,
gyromagnetic factor
n
/(2) (in units of MHz/T),
n
/(2) (in units of
10
4
K/T), natural abundance (in %), and electric quadrupole moment
Q (in units of 10
24
cm
2
).
Isotope I
n
/2
n
/2 Natural abundance Q
69
Ga 3/2 10.218 4.90 60.4 0.178
71
Ga 3/2 12.982 6.23 39.6 0.112
75
As 3/2 7.2919 3.50 100 0.3
ponential. In this limit, the average energy of a single nucleus writes as
E)
1
4
m=3/2

m=3/2
m
n
_
1
m
n
k
B
T
_
, (A.2)
and the Schottky nuclear specic heat per nucleus is given by
C
0
=
5
4

2
n

2
k
B
_
B
T
_
2
. (A.3)
For a single
69
Ga atom we obtain
69
C
0
= 0.3963
10
29
(B[T])
2
(T[K])
2
J/K. Averaging over the two Ga isotopes gives
Ga
C
0
= 1.25 [
69
C
0
]. The specic heat of a single As atom can be
expressed as
As
C
0
= 0.51 [
69
C
0
]. Consequently, provided that
n
k
B
T,
the Schottky nuclear heat capacity of Ga and As atoms in the QWs (C
QW
n
)
is estimated at ([
Ga
C
0
] +[
As
C
0
]) times the total number of GaAs molecules.
Appendix B
OPNMR Observations
Sensitivity poses a persistent challenge to nuclear magnetic resonance ex-
periments in the quantum Hall eect (QHE) regime, so spectroscopists have
resorted to schemes such as optical pumping in order to enhance the measur-
able signals. We summarize here some salient aspects of optically pumped
nuclear magnetic resonance (OPNMR) measurements in the QHE regime,
pioneered by Barrett and collaborators at Bell Laboratories and Yale Univer-
sity [7, 55, 56, 64, 103]. We start with a cursory outline of the experimental
technique.
The samples used in OPNMR experiments are electron-doped multiple-
quantum well (QW) GaAs/Al
0.1
Ga
0.9
As heterostructures. The NMR signal
is increased by optical pumping as follows. Polarized electrons are excited
in the conduction band by illuminating the sample with polarized light and
the strong hyperne coupling ensures the transfer of this polarization to
the nuclei. Since the incident photons have an energy tuned to the GaAs
bandgap only the nuclei located in the GaAs QWs will be polarized by this
process
1
. The bandgap of the barriers is larger than photons energy and
hence, there are no photo-excited electrons in Al
0.1
Ga
0.9
As regions. This op-
tical pumping strongly enhances the nuclear spin polarization (by as much
as a factor of 100), allowing NMR signals of the GaAs QWs to be inves-
tigated by direct, radio-frequency detection. Unfortunately, together with
this advantage one has the disadvantage that OPNMR signals exhibit quite
complicate dependences on the pumping wavelength, light polarization, and
power. Even though the initial excitation process is well understood, the
1
Note that, due to the local nature of the electron-nucleus interaction, the QWs nuclear
polarization prole in OPNMR measurements has a similar prole to the electron charge
density.
195
196 Appendix B: OPNMR Observations
recombination process, i.e., the relaxation of the electronic spins, is much
more complicated and may have an eect on the nuclear spin system of GaAs
QWs. Note also that in OPNMR experiments the two-dimensional electron
system (2DES) is observed while nuclei are strongly polarized, well beyond
their small equilibrium value. OPNMR measurements were carried out with
the timing sequence SAT-
L
-
D
-DET, where SAT represents a train of /2-
pulses which sets to zero the nuclear magnetization of the nuclear isotope of
interest,
L
is a period during which the sample is optically excited,
D
is a
period of no optical excitation, and DET represents the direct detection of
the free induction decay.
Table B.1:
71
Ga nuclear spin-lattice relaxation rate T
1
1
as a function of
temperature T and Landau level lling factor , as seen by the OPNMR
(reproduced with permission after Ref. [7]). OPNMR measurements found
minima in T
1
1
at 1 and 2/3 indicating energy gaps for electronic
excitations in both integer and fractional QHE states. The T-independent
T
1
1
at intermediate -values (2/3 1) suggest a manifold of low-lying
electronic states with mixed spin polarization.
= 1.01 = 0.88 = 0.66
T
1
1
(T = 4.2 K) [s
1
] 0.008 0.04 0.023
T
1
1
(T = 2.1 K) [s
1
] 0.0008 0.05 0.015
In Chapter 4, we guide the discussion of our NMR data on the basis of
the following four OPNMR observations. First, OPNMR experiments have
provided perspicuous evidence for the existence of nite-size Skyrmions near
Landau level lling factor = 1. More precisely, it was found in these
experiments that the low-temperature (T) electron spin polarization T()
drops precipitously on either side of = 1, which is evidence that the charged
excitations of the = 1 QHE ground state invoke many spin reversals.
OPNMR data revealed that the low-T electron spin polarization approaches
the full polarization (T = 1) at = 1, thereby providing additional credence
to the Skyrmion picture at = 1 [Fig. (4.12), panels (a) and (b)]. Of further
interest to us is the fact that the electron spin polarization peak at = 1 is
readily observed already at T = 4.2 K.
Secondly, the nuclear spin-lattice relaxation rate T
1
1
(T, ) of the
71
Ga
nuclei located in the GaAs QWs has been measured for 1.5 K T 4.2 K
Appendix B: OPNMR Observations 197
in the vicinity of = 1 [Fig. (4.12)(c)]. These OPNMR measurements
[Table (B.1)] strongly support the Skyrmion-mediated nuclear spin-lattice
relaxation as the underlying mechanism responsible for -dependence of T
1
1
near = 1 in disorder-free 2DESs.
The third observation merely records the fact that OPNMR experiments
studied the electron spin polarization of the 2DES around = 1/3. For ex-
ample, OPNMR measurements revealed that the T-saturated T() drops on
either side of = 1/3. The observed depolarization is quite small, consis-
tent with an average of 0.1 spin ips per quasihole (or quasiparticle). More
important, the low-T (T = 1) limit of OPNMR Knight shift (K
S
) data at
= 1/3, measured for several samples, was successfully used to determine
the eective hyperne coupling constant A
c
= (4.5 0.2) 10
13
cm
3
/s,
given by the relationship K
S
= A
c
Tn/w. Here w is the QW width and n is
the electron density.
Finally, OPNMR experiments also addressed the dynamics of the 2DES
by observing cross overs from the motionally narrowed regime to the
frozen regime
2
. This evolution is reminiscent of the behavior seen in
NMR studies where the motion of the nuclei freezes out [1]. In particular,
OPNMR spectra provided evidence that spin-reversed charged excitations of
the = 1/3 QHE ground state are localized at low temperatures (T 0.5 K)
over the NMR time scale of about 40 s. Information about the electron dy-
namics has been drawn from the T-dependence of the NMR line width. This
fourth observation is summarized, for convenience, in Table (B.2). When the
2DES crosses from the motionally-narrowed to the frozen regime, the width
of the QWs resonance increases dramatically. Qualitatively similar trends
were recently uncovered in low-T OPNMR measurements near = 1 [56].
2
What are the motionally narrowed and frozen regimes? The NMR spectrum can
have dierent line shapes, depending on the dynamics of the 2DES. An inhomogeneous
electron spin polarization will tend to broaden the NMR spectrum. Lets suppose that the
density of the electronic system is uniform, but the sample is divided into two regions, with
dierent electron spin polarizations. If the dynamic time of these regions is much slower
than the experimental NMR time scale, then we shall obtain a broadened NMR spectrum
(eventually composed of two peaks). However, if the dynamic time of these regions is much
faster than the experimental time scale, then the nuclei will feel the average polarization
of the two regions, and the NMR spectrum will tend to show a single (narrow) peak. The
former is the so-called frozen regime and the latter is the motionally-narrowed regime.
198 Appendix B: OPNMR Observations
Table B.2: The percentage increase of the QWs line width near = 1/3
observed in OPNMR experiments, relative to the value of 5.2 kHz measured
at T = 1.5 K and = 1/3 (reproduced with permission after Ref. [55]).

T [K]
1.5 0.9 0.7 0.5 0.3
0.33 0% 4% 4% 3% 5%
0.29 2% 12% 20% 36% 32%
0.27 12% 21% 45% 69% 53%
It is probably worth adding here that Sinova et al. [93] addressed the
question of how the average T is related to the NMR line shape. In the
motionally-narrowed regime, where T could be considered homogeneous, the
position of QWs line in the NMR spectrum (what is usually understood
by the Knight shift) could be considered as a direct measurement of T.
However, in the frozen regime, the QWs peak in the NMR spectrum is no
longer a good measure of T. Instead, in order to avoid dynamic eects, one
should measure the rst momentum of the QWs line shape which will give
T up to a proportionality constant.
Bibliography
[1] Abragam A., Principles of Nuclear Magnetism, (Oxford University
Press, New York, 1961).
[2] Aifer E.H., B.B. Goldberg, and D.A. Broido, Evidence of Skyrmion
Excitations about = 1 in n-Modulation-Doped Single Quantum Wells
by Interband Optical Transmission, Phys. Rev. Lett. 76, 680 (1996).
[3] Ando T. and Y. Uemura, Theory of Oscillatory g Factor in a MOS
Inversion Layer under Strong Magnetic Fields, J. Phys. Soc. Japan 37,
1044 (1974).
[4] Andraka B. and Y. Takano, Simultaneous Measurements of Heat Ca-
pacity and Spin-Lattice Relaxation Time in High Magnetic Field at Low
Temperature, Rev. Sci. Instrum. 67, 4256 (1996).
[5] Antoniou D. and A.H. MacDonald, Nuclear-Spin Lattice Relaxation and
Spin-Wave Collective Modes in a Disordered Two-Dimensional Electron
Gas, Phys. Rev. B 43, 11686 (1991).
[6] Bachmann R., F.J. DiSalvo Jr., T.H. Geballe, R.L. Greene, R.E.
Howard, C.N. King, H.C. Kirsch, K.N. Lee, R.E. Schwall, H.-U.
Thomas, and R.B. Zubeck, Heat Capacity Measurements on Small Sam-
ples at Low Temperatures, Rev. Sci. Instrum. 43, 205 (1972).
[7] Barrett S.E., G. Dabbagh, L.N. Pfeier, K.W. West, and R. Tycko,
Optically Pumped NMR Evidence for Finite-Size Skyrmions in GaAs
Quantum Wells near Landau Level Filling = 1, Phys. Rev. Lett. 74,
5112 (1995).
[8] Bayot V., M.B. Santos, and M. Shayegan, Oscillations in the Ther-
mal Conductivity of a GaAs/AlGaAs Heterostructure in the Fractional
Quantum Hall Regime, Phys. Rev. B 46, 7240 (1992).
199
200 BIBLIOGRAPHY
[9] Bayot V., E. Grivei, S. Melinte, M.B. Santos, and M. Shayegan, Giant
Low Temperature Heat Capacity of GaAs Quantum Wells near Landau
Level Filling = 1, Phys. Rev. Lett. 76, 4584 (1996).
[10] Bayot V., E. Grivei, J.-M. Beuken, S. Melinte, and M. Shayegan, Crit-
ical Behavior of Nuclear-Spin Diusion in GaAs/AlGaAs Heterostruc-
tures near Landau Level Filling = 1, Phys. Rev. Lett. 79, 1718 (1997).
[11] Berg A., M. Dobers, R.R. Gerhardts, and K. v. Klitzing, Magneto-
Quantum Oscillations of the Nuclear Spin-Lattice Relaxation near a
Two-Dimensional Electron Gas, Phys. Rev. Lett. 64, 2563 (1990).
[12] Blakemore J.S., Semiconducting and Other Major Properties of GaAs,
J. Appl. Phys. 53, R123 (1982).
[13] Boebinger G.S., H.L. Stormer, D.C. Tsui, A.M. Chang, J.C.M. Hwang,
A.Y. Cho, C.W. Tu, and G. Weimann, Activation Energies and Local-
ization in the Fractional QHE, Phys. Rev. B 36, 7919 (1987).
[14] Brey L., H.A. Fertig, R. Cote, and A.H. MacDonald, Skyrme Crystal
in a Two-Dimensional Electron Gas, Phys. Rev. Lett. 75, 2562 (1995).
[15] Carr H.Y. and E.M. Purcell, Eects of Diusion on Free Precession in
NMR Experiments, Phys. Rev. 94, 630 (1954).
[16] Chakraborty T. and F.C. Zhang, Role of the Reversed Spins in the
Correlated Ground State for the FQHE, Phys. Rev. B 29, 7032 (1984).
[17] Chakraborty T., P. Pietilainen, and F.C. Zhang, Elementary Exci-
tations in the Fractional QHE and the Spin-Reversed Quasiparticles,
Phys. Rev. Lett. 57, 130 (1986).
[18] Chakraborty T. and P. Pietilainen, The Quantum Hall Eects: Integral
and Fractional, (Springer-Verlag, 1995), 2nd ed.
[19] Chakraborty T., P. Pietilainen, and R. Shankar, Spin Polarizations
at and about the Lowest Filled Landau Level, Europhys. Lett. 38, 141
(1997).
[20] Cho H., J.B. Young, W. Kang, K.L. Campman, A.C. Gossard, M. Bich-
ler, and W. Wegscheider, Hysteresis and Spin Transitions in the Frac-
tional QHE, Phys. Rev. Lett. 81, 2522 (1998).
BIBLIOGRAPHY 201
[21] Clark R.G., S.R. Haynes, A.M. Suckling, J.R. Mallett, P.A. Wright, J.J.
Harris, and C.T. Foxon, Spin Congurations and Quasiparticle Frac-
tional Charge of Fractional QHE Ground States in the N = 0 Landau
Level, Phys. Rev. Lett. 62, 1536 (1989).
[22] Collan H.K., M. Krusius, and G.R. Pickett, Specic Heat of Antimony
and Bismuth between 0.03 and 0.8 K, Phys. Rev. B 1, 2888 (1970).
[23] Cooper N.R., Skyrmions in Quantum Hall Systems with Realistic Force
Laws, Phys. Rev. B 55, R1934 (1997).
[24] Cote R., A.H. MacDonald, L. Brey, H.A. Fertig, S.M. Girvin, and
H.T.C. Stoof, Collective Excitations, NMR, and Phase Transitions in
Skyrme Crystals, Phys. Rev. Lett. 78, 4825 (1997).
[25] Das Sarma S. and F. Stern, Single-Particle Relaxation Time versus
Scattering Time in an Impure Electron Gas, Phys. Rev. B 32, 8442
(1985).
[26] Das Sarma S. and X.C. Xie, Strong-Magnetic Field DOS in Weakly
Disordered 2DESs, Phys. Rev. Lett. 61, 738 (1988).
[27] Das Sarma S. and A. Pinczuk [editors], Perspectives in Quantum Hall
Eects, (Wileys, New York, 1997).
[28] Dobers M., K. v. Klitzing, J. Schneider, G. Weimann, and K. Ploog,
Electrical Detection of NMR in GaAs/AlGaAs Heterostructures, Phys.
Rev. Lett. 61, 1650 (1988).
[29] Du R.R., A.S. Yeh, H.L. Stormer, D.C. Tsui, L.N. Pfeier, and K.W.
West, Fractional QHE around = 3/2: Composite Fermions with a
Spin, Phys. Rev. Lett. 75, 3926 (1995).
[30] Eisenstein J.P., H.L. Stormer, V. Narayanamurti, A.Y. Cho, A.C. Gos-
sard, and C.W. Tu, Density of States and de Haas-van Alphen Eect
in 2DESs, Phys. Rev. Lett. 55, 875 (1985).
[31] Eisenstein J.P., H.L. Stormer, L.N. Pfeier, and K.W. West, Evidence
for a Spin Transition in the = 2/3 Fractional QHE, Phys. Rev. B 41,
7910 (1990).
[32] Engel L., S.W. Hwang, T. Sajoto, D.C. Tsui, and M. Shayegan, Frac-
tional QHE at = 2/3 and = 3/5 in Tilted Magnetic Fields, Phys.
Rev. B 45, 3418 (1992).
202 BIBLIOGRAPHY
[33] Eom J., H. Cho, W. Kang, K.L. Campman, A.C. Gossard, M. Bich-
ler, and W. Wegscheider, Quantum Hall Ferromagnetism in a 2DES,
Science 289, 2320 (2000).
[34] Fagaly R.L. and R.G. Bohn, A Modied Heat Pulse Method for Deter-
mining Heat Capacities at Low Temperatures, Rev. Sci. Instrum. 48,
1502 (1977).
[35] Fang F.F. and P.J. Stiles, Eects of a Tilted Magnetic Field on a 2DES,
Phys. Rev. 174, 823 (1968).
[36] Fertig H.A., L. Brey, R. Cote, A.H. MacDonald, A. Karlhede, and S.L.
Sondhi, Hartree-Fock Theory of Skyrmions in Quantum Hall Ferromag-
nets, Phys. Rev. B 55, 10671 (1997).
[37] Fowler A.B., F.F. Fang, W.E. Howard, and P.J. Stiles, Magneto-
oscillatory Conductance in Silicon Surfaces, Phys. Rev. Lett. 16, 901
(1966).
[38] Fukushima S.M. and S.B.W. Roeder, Experimental Pulse NMR: A Nuts
and Bolts Approach, (Addison-Wesley, Massachusetts, 1981), p. 380.
[39] Girvin S.M., in Topological Aspects of Low Dimensional Systems, Pro-
ceedings of the Les Houches Summer School of Theoretical Physics,
Session LXIX, 1998, edited by A. Comtet, T. Jolicoeur, S. Ouvry, and
F. David, (Springer-Verlag and EDP Sciences, 1999), p. 53.
[40] Gmelin, E., Modern Low-Temperature Calorimetry, Thermochimica
Acta 29, 1 (1979).
[41] Gopal E.S.R., Specic Heats at Low Temperatures, (Plenum Press, New
York, 1966).
[42] Gornik E., R. Lassnig, G. Strasser, H.L. Stormer, A.C. Gossard,
and W. Wiegmann, Specic Heat of Two-Dimensional Electrons in
GaAs/AlGaAs Multilayers, Phys. Rev. Lett. 54, 1820 (1985).
[43] Green A.G., Disorder and the Quantum Hall Ferromagnet, Phys. Rev.
B. 57, R9373 (1998).
[44] Green A.G., Quantum-Critical Dynamics of the Skyrmion Lattice,
Phys. Rev. B 61, R16299 (2000).
BIBLIOGRAPHY 203
[45] Guerrier D.J. and R.T. Harley, Calibration of Strain vs Nuclear
Quadrupole Splitting in III-IV Quantum Wells, Appl. Phys. Lett. 70,
1739 (1997).
[46] Hahn E.L., Spin Echoes, Phys. Rev. 80, 580 (1950).
[47] Halonen V., P. Pietilainen, and T. Chakraborty, Subband-Landau-Level
Coupling in the Fractional QHE in Tilted Magnetic Fields, Phys. Rev.
B 41, 10202 (1990).
[48] Halperin B.I., Theory of the Quantum Hall Conductance, Helv. Phys.
Acta 56, 75 (1983).
[49] Haug, R.J., K. von Klitzing, R.J. Nicholas, J.C. Maan, and G.
Weimann, Fractional QHE in Tilted Magnetic Fields, Phys. Rev. B
36, 4528 (1987).
[50] Haussmann R., Magnetization of Quantum Hall Systems, Phys. Rev. B
56, 9684 (1997).
[51] Heinonen O. [editor], Composite Fermions, (World Scientic, Singa-
pore, 1998).
[52] Iordanskii S.V., S.V. Meshkov, and I.D. Vagner, Nuclear-Spin Relax-
ation and Spin Excitons in a Two-Dimensional Electron Gas, Phys.
Rev. B 44, 6554 (1991).
[53] Kasner M. and A.H. MacDonald, Thermodynamics of Quantum Hall
Ferromagnets, Phys. Rev. Lett. 76, 3206 (1996).
[54] Kasner M., J.J. Palacios, and A.H. MacDonald, Quasiparticle Proper-
ties of Quantum Hall Ferromagnets, Phys. Rev. B 62, 2640 (2000).
[55] Khandelwal P., N.N. Kuzma, S.E. Barrett, L.N. Pfeier, and K.W.
West, Optically Pumped NMR Measurements of the Electron Spin Po-
larization in GaAs Quantum Wells near Landau Level Filling Factor
= 1/3, Phys. Rev. Lett. 81, 673 (1998).
[56] Khandelwal P., A.E. Dementyev, N.N. Kuzma, S.E. Barrett, L.N. Pfeif-
fer, and K.W. West, Spectroscopic Evidence for the Localization of Skyr-
mions near = 1 as T 0, Phys. Rev. Lett. 86, 5353 (2001).
[57] Kim J.H., I.D. Vagner, and L. Xing, Phonon-Assisted Mechanism for
Quantum Nuclear-Spin Relaxation, Phys. Rev. B 49, 16777 (1994).
204 BIBLIOGRAPHY
[58] Kivelson S., D.H. Lee, and S.C. Zhang, Global Phase Diagram in the
Quantum Hall Eect, Phys. Rev. B 46, 2223 (1992).
[59] Kosevich A.M., B.A. Ivanov, and A.S. Kovalev, Magnetic Solitons,
Phys. Rep. 194, 117 (1990).
[60] Kralik B., A.M. Rappe, and S.G. Louie, Variational Monte Carlo Cal-
culation of the Spin Gap in the = 1 Quantum Hall Liquid, Phys. Rev.
B 56, 4760 (1997).
[61] Kronm uller S., W. Dietsche, K. v. Klitzing, G. Denninger, W. Wegschei-
der, and M. Bichler, New Type of Electron Nuclear-Spin Interaction
from Resistively Detected NMR in the Fractional QHE Regime, Phys.
Rev. Lett. 82, 4070 (1999).
[62] Kukushkin I.V., K. v. Klitzing, and K. Eberl, Spin Polarization of Two-
Dimensional Electrons in Dierent Fractional States and around Filling
Factor = 1, Phys. Rev. B 55, 10607 (1997).
[63] Kukushkin I.V., K. v. Klitzing, and K. Eberl, Enhancement of the
Skyrmionic Excitations due to the Suppression of Zeeman Energy by
Optical Orientation of Nuclear Spins, Phys. Rev. B 60, 2554 (1999).
[64] Kuzma N.N., P. Khandelwal, S.E. Barrett, L.N. Pfeier, and K.W.
West, Ultraslow Electron Spin Dynamics in GaAs Quantum Wells
Probed by Optically Pumped NMR, Science 281, 686 (1998).
[65] Lanzillotto A.-M., M.B. Santos, and M. Shayegan, Silicon Migration
during the MBE of -Doped GaAs and Al
0.25
Ga
0.75
As, J. Vac. Sci.
Technol. A 8, 2009 (1990).
[66] Laughlin R.B., Quantized Hall Conductivity in Two Dimensions, Phys.
Rev. B 23, 5632 (1981).
[67] Laughlin R.B., Anomalous Quantum Hall Eect: An Incompressible
Quantum Fluid with Fractionally Charged Excitations, Phys. Rev. Lett.
50, 1395 (1983).
[68] Li Q., X.C. Xie, and S. Das Sarma, Calculated Heat Capacity and Mag-
netization of 2DESs, Phys. Rev. B 40, 1381 (1989).
[69] Lounasmaa O.V., Experimental Principles and Methods below 1 K,
(Academic Press, London, 1974), p. 277.
BIBLIOGRAPHY 205
[70] MacDonald A.H., H.C.A. Oji, and K.L. Liu, Thermodynamic Properties
of an Interacting Two-Dimensional Electron Gas in a Strong Magnetic
Field, Phys. Rev. B 34, 2681 (1986).
[71] MacDonald A.H. and J.J. Palacios, Magnons and Skyrmions in Frac-
tional Hall Ferromagnets, Phys. Rev. B 58, R10171 (1998).
[72] Malinowski A. and R.T. Harley, Anisotropy of the Electron g Factor in
Lattice-Matched and Strained-Layer III-IV Quantum Wells, Phys. Rev.
B 62, 2051 (2000).
[73] Manfra M.J., E.H. Aifer, B.B. Goldberg, D.A. Broido, L.N. Pfeier,
and K.W. West, Temperature Dependence of the Spin Polarization of a
Quantum Hall Ferromagnet, Phys. Rev. B 54, R17327 (1996).
[74] Manoharan H.C., New Particles and Phases in Reduced-Dimensional
Systems, Ph.D. Thesis, Princeton University (1998).
[75] Maude D.K., M. Potemski, J.C. Portal, H. Henini, L. Eaves, G. Hill,
and M.A. Pate, Spin Excitations of a Two-Dimensional Electron Gas
in the Limit of Vanishing Lande g Factor, Phys. Rev. Lett. 77, 4604
(1996).
[76] Melik-Alaverdian V., N.E. Bonesteel, and G. Ortiz, Skyrmions Physics
beyond the Lowest Landau-Level Approximation, Phys. Rev. B 60,
R8501 (1999).
[77] Melinte S., E. Grivei, V. Bayot, and M. Shayegan, Heat Capacity Evi-
dence for the Suppression of Skyrmions at Large Zeeman Energy, Phys.
Rev. Lett. 82, 2764 (1999).
[78] Melinte S., N. Freytag, M. Horvatic, C. Berthier, L.-P. Levy, V. Bayot,
and M. Shayegan, NMR Determination of Two-Dimensional Electron
Spin Polarization at = 1/2, Phys. Rev. Lett. 84, 354 (2000).
[79] Melinte S., N. Freytag, M. Horvatic, C. Berthier, L.-P. Levy, V. Bayot,
and M. Shayegan, Spin Polarization of Two-Dimensional Electrons in
GaAs Quantum Wells around Landau Level Filling = 1 from NMR
Measurements of Gallium Nuclei, Phys. Rev. B 64, 085327 (2001).
[80] Moon K., H. Mori, K. Yang, S.M. Girvin, A. H. MacDonald, L. Zheng,
D. Yoshioka, and S.-C. Zhang, Spontaneous Interlayer Coherence in
Double-Layer Quantum Hall Eect Systems: Charged Vortices and
Kosterlitz-Thouless Phase Transitions, Phys. Rev. B 51, 5138 (1995).
206 BIBLIOGRAPHY
[81] Morin F.J. and J.P. Maita, Specic Heats of Transition Metal Super-
conductors, Phys. Rev. 129, 1115 (1963).
[82] Nederveen A.J. and Yuli V. Nazarov, Skyrmions in Disordered Het-
erostructures, Phys. Rev. Lett. 82, 406 (1999).
[83] Nicholas R.J., R.J. Haug, K. von Klitzing, and G. Weimann, Exchange
Enhancement of the Spin Splitting in a GaAs/AlGaAs Heterojunction,
Phys. Rev. B 37, 1294 (1988).
[84] Paget D., Optical Detection in High-Purity GaAs: Direct Study of the
Relaxation of Nuclei close to Shallow Donors, Phys. Rev. B 25, 4444
(1982).
[85] Park K. and J.K. Jain, Phase Diagram of the Spin Polarization of
Composite Fermions and a New Eective Mass, Phys. Rev. Lett. 80,
4237 (1998).
[86] Prange R.E. and S.M. Girvin [editors], The Quantum Hall Eect,
(Springer-Verlag, New York, 1987).
[87] Read N. and S. Sachdev, Continuum Quantum Ferromagnets at Finite
Temperature and the QHE, Phys. Rev. Lett. 75, 3509 (1995).
[88] Richardson S.L., M.L. Cohen, S.G. Louie, and J.R. Chelikowsky, Elec-
tron Charge Densities at Conduction-Band Edges of Semiconductors,
Phys. Rev. B 33, 1177 (1986).
[89] Schmeller A., J.P. Eisenstein, L.N. Pfeier, and K.W. West, Evidence
for Skyrmions and Single Spin Flips in the Integer QHE, Phys. Rev.
Lett. 75, 4290 (1995).
[90] Shayegan M., J.K. Wang, M.B. Santos, T. Sajoto, and B.B. Goldberg,
Fractional QHE in a High-Mobility GaAs/AlGaAs Multiple-Quantum-
Well Heterostructure, Appl. Phys. Lett. 54, 27 (1989).
[91] Shepherd J.P., Analysis of the Lumped
2
Eect in Relaxation
Calorimetry, Rev. Sci. Instrum. 56, 273 (1985).
[92] Shukla S.P., M. Shayegan, S.R. Parihar, S.A. Lyon, N.R. Cooper, and
A.A. Kiselev, Large Skyrmions in an Al
0.13
Ga
0.87
As Quantum Well,
Phys. Rev. B 61, 4469 (2000).
BIBLIOGRAPHY 207
[93] Sinova J., S.M. Girvin, T. Jungwirth, and K. Moon, Skyrmion Dynam-
ics and NMR Line Shapes in Quantum Hall Ferromagnets, Phys. Rev.
B 61, 2749 (2000).
[94] Slichter C.P., Principles of Magnetic Resonance, (Springer-Verlag,
Berlin, 1991).
[95] Sondhi S.L., A. Karlhede, S.A. Kivelson, and E.H. Rezayi, Skyrmions
and the Crossover from the Integer to Fractional QHE at Small Zeeman
Energies, Phys. Rev. B 47, 16419 (1993).
[96] Song Yi-Qiao, B.M. Goodson, K. Maranowski, and A.C. Gossard, Re-
duction of Spin Polarization near Landau Level Filling Factor = 3 in
GaAs/AlGaAs Quantum Wells, Phys. Rev. Lett. 82, 2768 (1999).
[97] Stone M., Magnus Force on Skyrmions in Ferromagnets and Quantum
Hall Eect Systems, Phys. Rev. B 53, 16573 (1996).
[98] Sullivan P. and G. Seidel, Steady-State, ac-Temperature Calorimetry,
Phys. Rev. 173, 679 (1968).
[99] Timm C., S.M. Girvin, P. Henelius, and A. W. Sandvik, 1/N Expansion
for Two-Dimensional Quantum Ferromagnets, Phys. Rev. B 58, 1464
(1998).
[100] Timm C., S.M. Girvin, and H.A. Fertig, Skyrmion Lattice Melting in
the Quantum Hall System, Phys. Rev. B 58, 10634 (1998).
[101] Townes C.H., C. Herring, and W.D. Knight, The Eect of Electronic
Paramagnetism on NMR Frequencies in Metals, Phys. Rev. 77, 852(L)
(1950).
[102] Tsui D.C., H.L. Stormer, and A.C. Gossard, Two-Dimensional Mag-
netotransport in the Extreme Quantum Limit, Phys. Rev. Lett. 48, 1559
(1982).
[103] Tycko R., S.E. Barrett, G. Dabbagh, L.N. Pfeier, and K.W. West,
Electronic States in GaAs Quantum Wells Probed by Optically Pumped
NMR, Science 268, 1460 (1995).
[104] Usher A., R.J. Nicholas, J.J. Harris, and C.T. Foxon, Observation of
Magnetic Excitons and Spin Waves in Activation Studies of a Two-
Dimensional Electron Gas, Phys. Rev. B 41, 1129 (1990).
208 BIBLIOGRAPHY
[105] Vagner I.D. and T. Maniv, Nucler Spin-Lattice Relaxation: A Micro-
scopic Local Probe for Systems Exhibiting the QHE, Phys. Rev. Lett.
61, 1400 (1988).
[106] van der Pauw, L.J., A Method of Measuring Specic Resistivity and
Hall Eect of Discs of Arbitrary Shape, Philips Research Reports, 13,
1 (1958).
[107] Vitkalov S.A., C.R. Bowers, J.A. Simmons, and J.L. Reno, ESR Detec-
tion of Optical Dynamic Nuclear Polarization in GaAs/AlGaAs Quan-
tum Wells at Unity Filling Factor in the QHE, Phys. Rev. B 61, 5447
(2000).
[108] von Klitzing K., G. Dorda, and M. Pepper, New Method for High Accu-
racy Determination of the Fine-Structure Constant based on Quantized
Hall Resistance, Phys. Rev. Lett. 45, 494 (1980).
[109] Wang J.K., J.H. Campbell, D.C. Tsui, and A.Y. Cho, Heat Capac-
ity of the Two-Dimensional Electron Gas in GaAs/AlGaAs Multiple-
Quantum-Well Heterostructures, Phys. Rev. B 38, 6174 (1988).
[110] Wang J.K., ac Calorimetric Study of the Heat Capacity of Two-
Dimensional Electrons in Magnetic Field, Ph.D. Thesis, Princeton Uni-
versity (1990).
[111] Wang J.K., D.C. Tsui, M.B. Santos, and M. Shayegan, Heat Capac-
ity Study of Two-Dimensional Electrons in GaAs/AlGaAs Multiple-
Quantum-Well Heterostructures in High Magnetic Fields: Spin-Split
Landau Levels, Phys. Rev. B 45, 4384 (1992).
[112] Wiegers S.A.J, M. Specht, L.-P. Levy, M.Y. Symmons, D.A. Ritchie,
A. Cavanna, B. Etienne, G. Martinez, and P. Wyder, Magnetization
and Energy Gaps of a High-Mobility 2D Electron Gas in the Quantum
Limit, Phys. Rev. Lett. 79, 3238 (1997).
[113] Winter J., Magnetic Resonance in Metals, (Clarendon Press, Oxford,
1971).
[114] Zawadski W. and R. Lassnig, Specic Heat and Magneto-Thermal Os-
cillations of 2DESs in a Magnetic Field, Solid State Commun. 50, 537
(1984).
[115] Ziman J.M., Principles of the Theory of Solids, (Cambridge University
Press, Cambridge, 1972), p. 353.

You might also like